05.07.2013 Views

MAGNETISM ELECTRON TRANSPORT MAGNETORESISTIVE LANTHANUM CALCIUM MANGANITE

MAGNETISM ELECTRON TRANSPORT MAGNETORESISTIVE LANTHANUM CALCIUM MANGANITE

MAGNETISM ELECTRON TRANSPORT MAGNETORESISTIVE LANTHANUM CALCIUM MANGANITE

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>MAGNETISM</strong> AND <strong>ELECTRON</strong> <strong>TRANSPORT</strong> IN<br />

<strong>MAGNETORESISTIVE</strong> <strong>LANTHANUM</strong> <strong>CALCIUM</strong><br />

<strong>MANGANITE</strong><br />

A DISSERTATION<br />

SUBMITTED TO THE DEPARTMENT OF APPLIED PHYSICS<br />

AND THE COMMITTEE ON GRADUATE STUDIES<br />

OF STANFORD UNIVERSITY<br />

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS<br />

FOR THE DEGREE OF<br />

DOCTOR OF PHILOSOPHY<br />

G. Jeffrey Snyder<br />

June 1997


© Copyright by G. Jeffrey Snyder 1997<br />

All Rights Reserved<br />

ii


I certify that I have read this dissertation and that in my<br />

opinion it is fully adequate, in scope and in quality, as a<br />

dissertation for the degree of Doctor of Philosophy.<br />

_______________________________________<br />

Theodore H. Geballe<br />

(Principal Adviser)<br />

I certify that I have read this dissertation and that in my<br />

opinion it is fully adequate, in scope and in quality, as a<br />

dissertation for the degree of Doctor of Philosophy.<br />

_______________________________________<br />

Malcolm R. Beasley<br />

I certify that I have read this dissertation and that in my<br />

opinion it is fully adequate, in scope and in quality, as a<br />

dissertation for the degree of Doctor of Philosophy.<br />

_______________________________________<br />

Robert L. White<br />

(Materials Science Department)<br />

Approved for the University Committee on Graduate<br />

Studies:<br />

_______________________________________<br />

iii


Abstract<br />

It is the goal of this thesis to understand the physical properties associated<br />

with the large negative magnetoresistance found in lanthanum calcium<br />

manganite. Such large magnetoresistances have been reported that this<br />

material is being considered for use as a magnetic field sensor. However,<br />

there are many variables such as temperature, magnetic field, chemical<br />

composition and processing that greatly influence the magnitude of the<br />

magnetoresistance. After introducing the problem in Chapter 1, Chapters 2<br />

and 3 describe the materials synthesis and physical property measurements<br />

used in this work. In Chapter 4, the intrinsic magnetic and electron transport<br />

properties of lanthanum calcium manganite are distinguished from those<br />

that depend largely on the chemical synthesis and processing. Chemical<br />

substitution of lanthanum by gadolinium, discussed in Chapter 5, not only<br />

induces ferrimagnetism, but also dramatically alters the electron transport<br />

because of slight structural changes. The physical mechanisms and empirical<br />

relationships found among the resistivity, magnetoresistance and magnetism<br />

in Chapters 3 and 4 are studied in greater depth in Chapters 6 and 7 and<br />

compared with theoretical predictions. This analysis provides a useful<br />

method for predicting the magnetoresistance as a function of temperature,<br />

magnetic field and transition temperature. The related perovskite, strontium<br />

ruthenate, proves to be a model compound for the study of metallic<br />

ferromagnets. The results of this work is presented in two appendices, and<br />

compared with the manganite results throughout the text.<br />

v


Preface<br />

Looking back at the many years at Stanford, there are many people I would<br />

like to thank for helping me along the long, windy path to a Ph.D. thesis. My<br />

dad, papa Schneider, Dr. Demin, and Frank DiSalvo deserve the credit for<br />

getting me interested in science: chemistry, engineering, materials science and<br />

physics.<br />

The graduate first year classes at Stanford would have been far too<br />

unbearable without the support of my first year commiserators Weber, Jim<br />

and Shelly. My time at the Max Planck Institute in Stuttgart, Germany could<br />

not have been more productive or pleasant thanks to Prof. Dr. Arndt Simon,<br />

the whole Abteilung Simon and foreign student ghetto especially Paul Rauch,<br />

Thomas Braun and Chris Ewels.<br />

Having nothing to do with superconductivity, much of my work at<br />

Stanford was outside the KGB headquarters in Ginzton Lab. The materials<br />

synthesis for this project was done at the Center for Materials Research in the<br />

McCullough building. Bob Feigelson and his group deserve a special thanks<br />

for advice and use of equipment, such as the laser heated crystal growth<br />

apparatus. Some of the crystal samples used in this thesis were grown by<br />

Vlad Beffa, a Stanford undergraduate working as a CMR summer student. I<br />

would also like to thank the CMR support staff: Tracy Tingle with the SEM<br />

microprobe, Glen and Waldo for keeping up the x-ray facility, Ann Marshall<br />

for TEM studies, and Thomas Carlson and Mark Gibson for knowing how to<br />

get everything done in McCullough. I would especially like to thank Bob<br />

White, Shan Wang and their students for teaching me about magnetic and<br />

magnetoresistive materials in their group meetings.<br />

Thanks also to the KGB group, especially Ted and Mac for helping me get<br />

started. K. A. Moler performed most of the experiment and much of the<br />

analysis of the heat capacity experiment in Appendix B. Lior Klein was the<br />

vii


viii<br />

inspiration behind the irreversibility line plot in section 3.2.2.2.9. Thanks also<br />

to Khiem, Steve, Daniel & Jenny, Laurent and the rest of the curry club for<br />

great conversations at lunch.<br />

Much of this work was done in collaboration with Hewlett-Packard in Palo<br />

Alto. The MOCVD films of the manganites were made by Ron Hiskes and<br />

Steve DiCarolis. The XAFS studies in Chapter 5 were performed at the<br />

Stanford Linear Accelerator by Corwin Booth and Bud Bridges from U. C.<br />

Santa Cruz.<br />

This work could not have been completed without the support of friends,<br />

housemates, parents but particularly Sossina Ñ thank you very much.<br />

Funding for this research was generously provided by The Fanny and John<br />

Hertz Foundation, the Air Force Office of Scientific Research, and the<br />

Stanford Center for Materials Research under the NSF-MRL program


Table of Contents<br />

Abstract v<br />

Preface vii<br />

Table of Contents i x<br />

List of Figures xiii<br />

List of Tables xvii<br />

1. INTRODUCTION 1<br />

1.1 Motivation 2<br />

1.2 AMnO 3<br />

1.3 Double Exchange 5<br />

2. MATERIALS SYNTHESIS AND CHARACTERIZATION 9<br />

2.1 Sample Preparation 9<br />

2.1.1 Bulk Polycrystalline Samples 9<br />

2.1.2 Single Crystals 10<br />

2.1.2.1 Flux Growth 11<br />

2.1.2.2 Float Zone 12<br />

2.1.2.3 Thin Films 13<br />

2.1.3 Reactive Samples 13<br />

2.2 Characterization 1 4<br />

2.2.1 Elemental Analysis 14<br />

2.2.2 Structural Analysis 14<br />

2.2.2.1 Neutron diffraction 15<br />

2.2.2.2 X-ray diffraction 15<br />

2.2.2.2.1 Powder X-ray diffraction 16<br />

2.2.2.2.2 Single crystal and films 17<br />

2.2.2.3 X-ray Absorption Fine Structure 17<br />

3. <strong>ELECTRON</strong>IC AND MAGNETIC MEASUREMENTS 1 9<br />

3.1 Transport Properties 1 9<br />

3.1.1.1 Ohm’s Law 19<br />

3.1.1.2 Magnetoresistance 20<br />

3.1.1.3 Drift velocity, mobility, relaxation time and mean free path 21<br />

3.1.1.4 Hall effects 22<br />

3.1.2 Measurement 22<br />

3.1.2.1 Linearity 23<br />

3.1.2.2 Geometry 24<br />

ix<br />

3


x<br />

3.1.2.3 Contacts 26<br />

3.1.2.4 Reproducibility 27<br />

3.1.2.5 Apparatus 28<br />

3.1.3 Analysis 29<br />

3.1.3.1 Metals 30<br />

3.1.3.1.1 Impurity scattering 30<br />

3.1.3.1.2 Electron-electron scattering 31<br />

3.1.3.1.3 Electron-phonon scattering 31<br />

3.1.3.2 insulators/semiconductors 31<br />

3.1.3.2.1 Band insulators/semiconductors 32<br />

3.1.3.2.2 Polarons 33<br />

3.1.3.2.3 Diffusive Conductivity 35<br />

3.1.3.2.4 Variable range Hopping 35<br />

3.1.3.3 Poor Metals / Heavily doped semiconductors 37<br />

3.1.3.4 Phase transitions 38<br />

3.2 Magnetism 3 9<br />

3.2.1 Measurement 39<br />

3.2.1.1 Apparatus 39<br />

3.2.2 Analysis 43<br />

3.2.2.1 Diamagnetism and Paramagnetism 44<br />

3.2.2.1.1 Larmor diamagnetism 44<br />

3.2.2.1.2 Conduction electron diamagnetism 46<br />

3.2.2.1.3 Pauli paramagnetism 46<br />

3.2.2.1.4 Curie paramagnetism 47<br />

3.2.2.2 Ferromagnetism 49<br />

3.2.2.2.1 Weiss molecular field model 50<br />

3.2.2.2.2 Itinerant electron Model 53<br />

3.2.2.2.3 Generalized Model 55<br />

3.2.2.2.4 Critical region 59<br />

3.2.2.2.5 Landau mean field theory 60<br />

3.2.2.2.6 Arrott Plot 61<br />

3.2.2.2.7 The Curie temperature 61<br />

3.2.2.2.8 Spin waves 63<br />

3.2.2.2.9 Irreversibility 66<br />

3.2.2.3 Antiferromagnetism 69<br />

3.2.2.4 Ferrimagnetism 71<br />

3.2.2.4.1 Mean field model for Gd0.67Ca0.33MnO3 72<br />

3.3 Heat Capacity 7 6<br />

3.3.1 Measurement 77<br />

3.3.1.1 Apparatus 78<br />

3.3.2 Analysis 78<br />

3.3.2.1 Electronic specific heat 78<br />

3.3.2.2 Phonon specific heat 79<br />

4. INTRINSIC ELECTRICAL <strong>TRANSPORT</strong> AND MAGNETIC<br />

PROPERTIES OF LA 0.67 CA 0.33 MNO 3 AND LA 0.67 SR 0.33 MNO 3 MOCVD<br />

THIN FILMS AND BULK MATERIAL 8 0<br />

4.1 Magnetism 8 1<br />

4.1.1 Low Temperature Excitations 83


4.2 Electronic Transport 8 5<br />

4.2.1 Low Temperature Resistivity 86<br />

4.2.1.1 Temperature independent term 89<br />

4.2.1.2 T2 dependent term 89<br />

4.2.1.3 Relationship to magnetism 91<br />

4.2.2 High Temperature resistivity 92<br />

4.2.3 Transport Near TC 93<br />

4.2.4 Hall effect 94<br />

4.2.5 Crystallographic Phase change 96<br />

4.2.6 Small Polarons 96<br />

4.2.7 Colossal Magnetoresistance 97<br />

4.2.8 Domain Boundary Magnetoresistance 99<br />

4.2.9 Low temperature magnetoresistance 99<br />

4.3 Conclusion 9 9<br />

5. LOCAL STRUCTURE, <strong>TRANSPORT</strong> AND RARE EARTH<br />

<strong>MAGNETISM</strong> IN THE FERRIMAGNETIC PEROVSKITE<br />

GD 0.67 CA 0.33 MNO 3<br />

xi<br />

101<br />

5.1 Ferrimagnetism 102<br />

5.1.1 Low temperature moment 103<br />

5.1.2 High temperature susceptibility 104<br />

5.1.3 Low temperature susceptibility 105<br />

5.1.4 Near T C magnetism 107<br />

5.1.5 Magnetism model 108<br />

5.1.5.1 Mean Field Model 108<br />

5.1.5.2 Canted antiferromagnetism 109<br />

5.1.5.3 Spin glass magnetism 110<br />

5.1.5.4 Related Compounds 111<br />

5.2 Electronic Transport 112<br />

5.2.1 Magnetoresistance 113<br />

5.2.2 Small Polaron Hopping 113<br />

5.2.3 Variable Range Hopping 114<br />

5.3 X-ray Absorption Fine Structure 114<br />

5.3.1 Relationship of structure to CMR 116<br />

5.4 Conclusion 116<br />

6. MAGNETOCONDUCTIVITY IN LA 0.67 CA 0.33 MNO 3<br />

118<br />

6.1 Anisotropic magnetoresistance 118


xii<br />

6.2 Magnetoresistance models 119<br />

6.2.1 General Model 120<br />

6.2.1.1 Magnetoconductivity model 121<br />

6.2.1.1.1 T > T C regime 123<br />

6.2.1.1.2 T < T C regime 124<br />

6.2.1.1.3 Anisotropic magnetoresistance 125<br />

6.2.1.1.4 Hall Effect 126<br />

6.3 Conclusion 127<br />

7. CRITICAL <strong>TRANSPORT</strong> AND MAGNETIZATION OF<br />

LA 0.67 CA 0.33 MNO 3<br />

128<br />

7.1 Magnetism near TC 130<br />

7.1.1 Spontaneous magnetization exponent 132<br />

7.1.2 Susceptibility exponent 134<br />

7.1.3 Positive nonlinear susceptibility 135<br />

7.1.4 Additional magnetic interaction 137<br />

7.2 Magnetoresistance 139<br />

7.2.1 Magnetoresistance scaling above TC 142<br />

7.2.2 Magnetoresistance scaling below TC 144<br />

7.2.3 Magnetoresistance scaling at TC 148<br />

7.2.4 Relation to low temperature magnetoresistance 149<br />

7.3 Conclusion 150<br />

Appendix A. Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

Appendix B. Magnetic Excitations and Specific Heat in SrRuO 3<br />

References 182<br />

152<br />

173


List of Figures<br />

xiii<br />

Figure 1-1 The Perovskite structure AMnO 3 where A is a mixture of rare earth and<br />

alkaline earth elements e.g. La 0.67Ca 0.33. 3<br />

Figure 1-2 Double exchange and the electronic structure of AMnO 3. 6<br />

Figure 3-1 Effect of changing the scan length for magnetization measurements. A<br />

YIG crystal is used. The three data reduction schemes are also compared. The<br />

results for the full scan algorithm with scan length less than 5 cm are off scale.41<br />

Figure 3-2 Magnetization of YIG sample as it is rotated along the field axis. 4 3<br />

Figure 3-3 Diamagnetic magnetic susceptibility of typical substrates. The increase<br />

in the susceptibility at low temperatures is due to paramagnetic impurities. 4 5<br />

Figure 3-4 Paramagnetic susceptibility and hysteresis loop of a paramagnetic Fe<br />

containing organometallic compound [78]. 4 8<br />

Figure 3-5 Effective paramagnetic moment of Fe in the organometallic compound<br />

[SC(CH 3) 2C(CH 3)NCH 2CH 2CH 2] 2N - FePF - 6 showing a spin transition [78]. 4 9<br />

Figure 3-6 Calculated inverse magnetic susceptibility of SrRuO 3 using the molecular<br />

field model. 50<br />

Figure 3-7 Mean field magnetization calculated in various fields for SrRuO 3 with T C =<br />

165K. Inset show the very small field dependence of the magnetization (forced<br />

magnetization) in this model. 5 2<br />

Figure 3-8 Energy spectrum of magnetic excitations. Spin wave excitations have a<br />

one-to-one dispersion relation while excitations in the Stoner continuum (shaded<br />

region) do not. The intensity of excitations in the Stoner continuum is strongest<br />

where the spin waves meet the continuum. 5 6<br />

Figure 3-9 The inflection T C measured for a SrRuO 3 pellet. For H < 1 Tesla the<br />

inflection T C is within 1 K of the Arrott T C = 163 K. At higher H the inflection<br />

T C increases by only a few degrees. 6 3<br />

Figure 3-10 Correction factor to the T 3/2 contribution of the magnetization in the spin<br />

wave theory due to a magnetic field H . 64<br />

Figure 3-11 Correction factor to the T 3/2 contribution of the heat capacity in the spin<br />

wave theory due to a magnetic field H . 65<br />

Figure 3-12 SrRuO 3 showing spin-glass like irreversibility of zero-field-cooled and<br />

field-cooled measurements in a small field. The field cooled curve may look<br />

saturated, but is actually less than 1/10 saturated at low temperatures. A small<br />

peak is observed in the zero-field-cooled measurement when the reversibility<br />

point is reached. 6 6<br />

Figure 3-13 Initial magnetization of SrRuO 3 pellet at 5 K, after cooling in zero field.<br />

The magnetization follows a “S” shaped curve providing an inflection point. 6 7<br />

Figure 3-14 Magnetic irreversibility line for polycrystalline SrRuO 3. Above the line<br />

the magnetization is reversible, below it is irreversible. The irreversibility<br />

exponent is about 1.5. 6 8<br />

Figure 3-15 Time dependent magnetization of SrRuO 3 pellet at 5 K. The field was<br />

increased from 0 to 100 Gauss. The magnetization follows a Log(time)<br />

dependence. 70<br />

Figure 3-16 Magnetic susceptibility of a Pt containing “CaRuO 3” crystal. The<br />

crystal was aligned with its 2-fold symmetric axis parallel to the applied field has<br />

a susceptibility characteristic of antiferromagnetic moments aligning parallel to<br />

the field, while the 3-fold axis appears to have moments perpendicular to the<br />

field. 71<br />

Figure 3-17 Temperature - tolerance factor phase diagram from reference [100], with<br />

the position of Gd 0.67Ca 0.33MnO 3 indicated. 7 3<br />

Figure 3-18 Calculated Arrott Plot for Gd 0.67Ca 0.33MnO 3 using the mean field model<br />

with T C = 83.3 K. 7 5


xiv<br />

Figure 3-19 High field differential susceptibility for Gd 0.67Ca 0.33MnO 3 calculated using<br />

the mean field model. The maximum is at 11.5 K which is near T Comp = 14.2 K in<br />

this model. 76<br />

Figure 4-1 Magnetization of La 0.67Sr 0.33MnO 3 polycrystalline pellet at 10kOe. Inset a,<br />

magnetization at 100Oe used to determine T C = 375K. Inset b, full hysteresis<br />

loop at 5 K. 81<br />

Figure 4-2 Magnetization of La 0.67Sr 0.33MnO 3 film (LSM1) on LaAlO 3 at 5kOe. Inset,<br />

full hysteresis loop at 5 K of film and (diamagnetic) substrate. 8 2<br />

Figure 4-3 Magnetization of La 0.67Ca 0.33MnO 3 film (LCM15) on LaAlO 3 at 5kOe.<br />

Inset, full hysteresis loop at 5 K of film and (diamagnetic) substrate. 8 3<br />

Figure 4-4 Magnetization of La 0.67(Ca/Sr) 0.33MnO 3 films and polycrystalline samples<br />

showing the T 2 dependence of the magnetization. Inset, same data as a function<br />

of T 3/2 for comparison. 8 4<br />

Figure 4-5 Comparison of the magnetization of La 0.67Sr 0.33MnO 3 with the T 3/2 term<br />

found at low temperatures, and various fits to the magnetization. 8 5<br />

Figure 4-6 Magnetoresistance of La 0.67Sr 0.33MnO 3 polycrystalline pellet and Film<br />

(LSM1). Inset, simultaneous magnetization and resistivity of the film at<br />

20Oersted, along with the magnetoresistance [R(H = 0 kOe)-R(H = 70 kOe)]. 8 6<br />

Figure 4-7 Magnetoresistance of La 0.67Ca 0.33MnO 3 film (LCM17). Inset, simultaneous<br />

magnetization and resistivity at 20Oersted, along with the magnetoresistance<br />

[R(H = 0kOe) - R( H = 70kOe)]. 8 7<br />

Figure 4-8 Low temperature resistivity (in zero field) of La 0.67(Sr/Ca) 0.33MnO 3 films<br />

(LSM1 and LCM10). Solid lines are the fit to R 0 + R 2T 2 + R 4.5T 4.5 up to 250K<br />

and 200K for LSM and LCM respectively. The dashed lines show the constant<br />

and T 2 terms of the best fit. 9 0<br />

Figure 4-9 High temperature resistivity (warming and cooling) of La 0.67Ca 0.33MnO 3 film<br />

(LCM17) and crystal in zero field. Inset a, same data with different abscissa to<br />

compare small polaron and semiconductor models. Inset b, DSC trace of<br />

polycrystalline La 0.67Ca 0.33MnO 3 showing the heat of the high temperature<br />

structural transition. 9 2<br />

Figure 4-10 High temperature resistivity (warming and cooling) of La 0.67Sr 0.33MnO 3<br />

film (LSM1) in zero field. Inset, same data displayed as in Figure 4-9. 9 3<br />

Figure 4-11 Resistance as a function of field La 0.67(Ca/Sr) 0.33MnO 3 films (LSM1 and<br />

LCM19) in the Hall effect configuration at 5 K. The Hall effect is calculated<br />

from the slope of the line indicated (see text). 9 5<br />

Figure 4-12 Colossal magnetoresistive La 0.67Ca 0.33MnO 3 film from [21, 127]. 9 7<br />

Figure 4-13 Simultaneous magnetization and resistivity in a magnetic field of<br />

La 0.67Sr 0.33MnO 3 polycrystalline pellet at 5 K. Data for both increasing and<br />

decreasing field are shown. Inset, Magnetoresistance of La 0.67Ca 0.33MnO 3 film<br />

(LCM10) at 5 K. 9 8<br />

Figure 5-1 Low temperature magnetization of Gd 0.67Ca 0.33MnO 3 measured in a 5 kOe<br />

field and zero field after cooling in a large field (remnant). 103<br />

Figure 5-2 Inverse magnetic susceptibility of bulk Gd 0.67Ca 0.33MnO 3. Solid line is the<br />

high temperature fit to χ = μ eff 2 /(8(T-Θ)) described in the text. 105<br />

Figure 5-3 Low temperature and high-field magnetic susceptibility, χ = (M (60 kOe)-<br />

M (40 kOe))/20kOe, of Gd 0.67Ca 0.33MnO 3 crystal. Inset, hysteresis loop at 5 K. 1 0 6<br />

Figure 5-4 Arrott plot of polycrystalline Gd 0.67Ca 0.33MnO 3 pellet. 107<br />

Figure 5-5 Magnetization and inverse magnetic susceptibility calculated for<br />

Gd 0.67Ca 0.33MnO 3 using the simplified mean field theory described in the text and T C<br />

= 83 K, T Comp = 17 K. The contribution to the magnetization of each sublattice is<br />

shown in dashed lines. 108<br />

Figure 5-6 High temperature resistivity during heating and cooling a Gd 0.67Ca 0.33MnO 3<br />

film, ln(ρ /T ) vs. 1/T . Inset a, comparison with ln(ρ ) vs. 1/T . Inset b,<br />

comparison with ln(ρ) vs. 1/T 1/4 . 111


Figure 5-7 Low temperature resistivity of Gd 0.67Ca 0.33MnO 3 crystal, ln(ρ) vs. 1/T 1/4 .<br />

Inset a, comparison with ln(ρ/T ) vs. 1/T . Solid lines show linear best fit to the<br />

data shown. Inset b, magnetoresistance of a film at 200 K and 300 K; solid line<br />

is the quadratic fit. 112<br />

Figure 5-8 Fourier transform of kχ(k) from (a) Mn K-edge and (b) Gd L III-edge data on<br />

Gd 0.67Ca 0.33MnO 3 . The solid lines are data collected at T = 69 K, while the<br />

triangles (Δ) are data collected at T = 40 K. Agreement between data above and<br />

below T C is well within the errors of the experiment. Transform ranges for the<br />

Gd edge data are from 3.5-12.5 Å -1 and Gaussian broadened by 0.3 Å -1 . Transform<br />

ranges for the Mn edge data are from 3.2-12.5 Å -1 and Gaussian broadened by 0.3<br />

Å -1 . 115<br />

Figure 6-1 High field (longitudinal) magnetoresistance above and below T C for<br />

La 0.67Ca 0.33MnO 3 film. The solid lines show the fit using the indicated equivalent<br />

circuit 123<br />

Figure 6-2 Low field magnetoresistance and magnetization (relative units) of<br />

La 0.67Ca 0.33MnO 3 at 0.9 T C. The sum of the longitudinal and transverse resistances<br />

minimizes the effect of the anisotropic magnetoresistance. 124<br />

Figure 6-3 Hall effect of La 0.67Ca 0.33MnO 3 below (fully magnetized data only) and<br />

above T C. 126<br />

Figure 7-1. M 2 vs. H /M plot for La 0.67Ca 0.33MnO 3 float zone crystal. A mean field<br />

ferromagnet has linear isotherms with a positive slope. The negative slope for T<br />

>T C indicates a faster than linear increase in M (inset) due to a highly unusual<br />

positive non-linear susceptibility χ 3. 130<br />

Figure 7-2. Data from Figure 7-1 (using the same symbols) scaled with β = 0.27 and<br />

γ = 0.90. According to the scaling hypothesis, all the T < T C data should lie on<br />

a single curve while the T > T C data should lie on a separate, single curve. 131<br />

Figure 7-3. Saturation Magnetization, M 0 as a function of temperature for<br />

La 0.67Ca 0.33MnO 3 crystal. At each temperature, the value shown is M extrapolated<br />

to H = 0 as given by the intercept in Figure 7-1.Solid line is fit to M 0(T ) ∝ (1 -<br />

T /T C) β with β = 0.30. 1 3 3<br />

Figure 7-4. Inverse magnetic susceptibility, 1/χ 0 as a function of temperature for<br />

La 0.67Ca 0.33MnO 3 crystal. At each temperature, the value shown is H/M<br />

extrapolated to H = 0 as given by the intercept in Figure 7-1. Solid line is fit to<br />

1/χ 0 ∝ (T /T C - 1) γ with γ = 0.7 and T C = 263K. 136<br />

Figure 7-5. Magnetization in a magnetic field for a La 0.67Ca 0.33MnO 3 polycrystalline<br />

pellet at 0.9 and 1.1 T C. The solid lines indicate the linear regions in each case.139<br />

Figure 7-6. Magnetoresistance of La 0.67Ca 0.33MnO 3 film compared with -M 2 of a pellet,<br />

both at 0.9 T C. The solid line for the magnetoresistance data shows the fit using<br />

the indicated equivalent circuit. The dashed line in the inset compares the<br />

exponential fit. 140<br />

Figure 7-7. Magnetoresistance of La 0.67Ca 0.33MnO 3 film compared with -M 2 of a pellet,<br />

both at 1.1 T C. The solid line for the magnetoresistance data shows the fit using<br />

the indicated equivalent circuit. 141<br />

Figure 7-8. Fitting parameters σ 0 and ρ ∞ for T > T C in a La 0.67Ca 0.33MnO 3 film. The<br />

temperature dependence of these two parameters reflect the insulating behavior of<br />

the material. 142<br />

Figure 7-9. Fitting parameter σ H2 as a function of temperature in a La 0.67Ca 0.33MnO 3<br />

film for T > T C. The temperature dependence of σ H2 and the square of the<br />

susceptibility are the same, indicating a relationship between the<br />

magnetoconductance and M 2 . 143<br />

Figure 7-10. Fitting parameters σ 0 and ρ ∞ for T < T C in a La 0.67Ca 0.33MnO 3 film. ρ ∞ is<br />

governed by the A + BT 2 terms in the resistivity while σ 0 diverges at T C. The<br />

inset shows σ 0 data fit with a (T C - T) 1.8 power law (dashed line), and σ ∝<br />

exp(M /M E) (solid line). The zero field resistivity ρ(H = 0) = ρ ∞ + 1/σ 0 is shown<br />

for comparison. 145<br />

xv


xvi<br />

Figure 7-11. Fitting parameter σ H as a function of temperature in a La 0.67Ca 0.33MnO 3<br />

film for T < T C. The solid line shows the best fit to the data using a critical<br />

exponent of 0.7. 146<br />

Figure 7-12. Magnetoresistance of La 0.67Ca 0.33MnO 3 film at 262 K ≈ T C. The solid line<br />

shows the fit (for the full data on a linear scale) using the indicated equivalent<br />

circuit. 149<br />

Figure A- 1. Magnetization at 5 K of SrRuO 3 single crystal along several<br />

crystallographic directions showing strong cubic but not uniaxial<br />

magnetocrystalline anisotropy. Inset shows the full hysteresis loop of the single<br />

crystal data along with that of a polycrystalline pellet for comparison. 155<br />

Figure A- 2. Arrott Plot of SrRuO 3 single crystal along easy [110] direction. Inset,<br />

critical isotherm (T = 163K ≈ T C) on a log scale fit to M δ ∝ H with δ = 4.2. 1 5 7<br />

Figure A- 3. Zero field magnetization M 0 of SrRuO 3 single crystal along easy [110]<br />

direction. Solid line shows the fit to M 0(T ) ∝ (1 - T /T C) β with β = 0.36. Inset<br />

showing the same data on a log plot. The critical exponent β appears to change<br />

from Heisenberg-like β = 0.39 near T C to Ising-like β = 0.32 as T decreases. 158<br />

Figure A- 4. Zero field inverse susceptibility 1/χ 0 of SrRuO 3 single crystal along<br />

easy [110] direction. Solid line shows the fit to 1/χ 0(T ) ∝ (1 - T /T C) γ with γ =<br />

1.17 and T C = 163.2 K. The inset shows the same data on a log plot. 159<br />

Figure A- 5. Scaled Arrott Plot of SrRuO 3 single crystal along easy [110] direction<br />

with β = 0.36 and γ =1.17. Symbols are the same as those used in Figure A- 2.160<br />

Figure A- 6. Magnetization as a function of temperature of SrRuO 3 single crystal<br />

along easy [110] direction. Inset shows the approximate T 2 dependence of the<br />

magnetization. 161<br />

Figure A- 7. Inverse magnetic susceptibility (1/χ = M /H ) at H = 10 kOe of<br />

polycrystalline SrRuO 3 compared to the single crystal data from Figure A- 4.<br />

The solid line is the straight-line fit with T C = 165K which demonstrates the<br />

slightly positive curvature of the data. 164<br />

Figure A- 8. Variation of the T 3/2 parameter in fitting the magnetization data of single<br />

crystal SrRuO 3 to M = M S (1 - AT 3/2 - BT 2 ) as the fitting range is increased. The<br />

upper inset shows the correlation of the A and B parameters. In the region where<br />

A is relatively stable (around T max = 60 K), A decreases as T max is lowered. The<br />

symbols are the same as those used in Figure A- 9. 166<br />

Figure A- 9. Variation of the T 2 parameter in fitting the magnetization data of single<br />

crystal SrRuO 3 to M = M S (1 - AT 3/2 - BT 2 ) as the fitting range is increased. In<br />

the region where B is relatively stable (around T max = 60 K), B increases as T max is<br />

lowered. 167<br />

Figure A- 10. Variation of the T 2 parameter in fitting the magnetization data of<br />

single crystal SrRuO 3 to M = M S (1 - BT 2 ) as the fitting range is increased. The<br />

parameter B for this fit is more stable and constant than that shown in Figure A-<br />

8. Inset, variation of Θ 2 in a magnetic field. 169<br />

Figure A- 11. Variation of A and B fitting parameters in the hypothetical case where<br />

the true magnetization is given by T 3/2 and T 5/2 terms. 170<br />

Figure B- 1 Heat capacity of SrRuO 3 cooled in zero field (zfc), in an 8 T magnetic<br />

field, and in zero field after being magnetized (rem). Inset, difference between the<br />

heat capacity measured after cooling in zero field with that in 8 T and the<br />

remnant magnetized state. 174<br />

Figure B- 2. Zero field heat capacity data fit with two free parameters, γ and β. Solid<br />

line, including the T 3/2 contribution expected theoretically for spin waves (see<br />

text). Dashed line, without any T 3/2 contribution. 175<br />

Figure B- 3. The linear term of the heat capacity γ as a function of magnetic field.<br />

Each circle is from a single c P,H datum between 4.3 and 5 K with phonons<br />

subtracted: γ(H ) = (c P,H(T) - βT 3 )/T with β = 0.191 mJ/mol·K 4 . γ(H ) for each<br />

square was determined by fitting 15-20 data points between 6 and 12 K to: c P,H(T)


xvii<br />

= γT + βT 3 . The triangles are calculated from the magnetization data of a single<br />

crystal. 176<br />

List of Tables<br />

Table 3-1 Theoretical 3-dimentional critical exponents for different models and<br />

selected experimental values [86, 87]. 6 0<br />

Table 4-1 Physical Properties of Polycrystalline Pellets 8 2<br />

Table 4-2 Magnetoresistance of Annealed Films. 8 9<br />

Table 5-1 Transition Temperatures for Gd 0.67Ca 0.33MnO 3. 107


<strong>MAGNETISM</strong> AND <strong>ELECTRON</strong> <strong>TRANSPORT</strong> IN<br />

1. Introduction<br />

<strong>MAGNETORESISTIVE</strong> LA 0.67CA 0.33MNO 3<br />

The development of new materials for technological applications has<br />

opened many doors to innovation in the 20 th century. New electronic and<br />

magnetic materials in particular have helped bring about the information<br />

revolution. Much of the progress is due to materials processing.<br />

Technological applications often have strict compositional and<br />

microstructural requirements for their materials. An integrated circuit for<br />

instance must have several compatible semiconductor, dielectric, and<br />

metallic materials with specific properties in precise locations.<br />

Improvements using well understood materials such as these are usually<br />

incremental.<br />

A risky but potentially more revolutionary method for advancing<br />

technologies is to find a different materials which have inherent properties<br />

superior to those currently in use. There are many known materials which<br />

need to be better understood before it would be clear that their use would be a<br />

significant advancement. In some cases a previously unknown class of<br />

compounds (such as the cuprate superconductors) may have to be discovered.<br />

It is also important to consider other aspects of the material, such as chemical<br />

and thermal stability, toxicity and availability.<br />

The study of new materials physics can have different emphasis. Many<br />

physicists are interested in new materials because they can be used to study a<br />

new physical phenomenon. An example of this is the study of heavy<br />

fermion metals and superconductors which have little potential application<br />

1


2 Chapter 1<br />

in themselves, but the physics learned from their study may be quite useful.<br />

Conversely, one can use physics to help understand new materials for<br />

potential applications. The physics may be well established but will give<br />

valuable insight into the uses and limitations of the material. The emphasis<br />

of this dissertation is on the latter: what physics can reveal about a material<br />

as opposed to what the material can tell you about physics.<br />

1.1 Motivation<br />

This dissertation has been motivated by the desire to understand the basic<br />

transport and magnetic properties and the physics behind them in metallic,<br />

ferromagnetic perovskite oxides. The manganites in particular show a wealth<br />

of complex properties. It has been useful to characterize these properties as<br />

either common to ferromagnetic metals in general or unique to the<br />

manganites. For this reason, the study of SrRuO 3 has been quite useful in<br />

understanding the properties of the manganites. For example, SrRuO 3 and<br />

the manganites shows similar magnetic critical behavior and low<br />

temperature magnetic excitations. It has also been advantageous to further<br />

classify the properties of the manganites as those which are intrinsic to the<br />

material and those affected by processing. The physics of the intrinsic<br />

properties are easier to study, while the extrinsic properties can be easily<br />

modified. Once the inherent properties of the material are understood,<br />

properties which depend strongly on processing can then be tuned for used in<br />

a device with particular characteristics.<br />

In this chapter some background on the manganites is presented, focusing<br />

in the recent interest in ÒColossal magnetoresistanceÓ (CMR). Chapter 2<br />

briefly summarizes the materials synthesis and characterization. In chapter 3<br />

the magnetism and transport experimental procedures are given as well as<br />

the pertinent analysis and theory. In chapter 4 the results of the intrinsic<br />

electrical and magnetic properties of La 0.67 Ca 0.33 MnO 3 are presented. In<br />

chapter 5 the ferrimagnetism and structure-transport correlations are shown


A<br />

Introduction 3<br />

Oxygen<br />

Mn<br />

Figure 1-1 The Perovskite structure AMnO 3 where A is a<br />

mixture of rare earth and alkaline earth elements e.g.<br />

La 0.67 Ca 0.33 .<br />

for the related compound Gd 0.67 Ca 0.33 MnO 3 . Chapter 6 introduces the<br />

magnetoconductivity analysis of the magnetotransport phenomena studied<br />

here. In Chapter 7 this analysis is used to examine the relationship bewteen<br />

the magnetization and the magnetoresistance.<br />

1.2 AMnO 3<br />

The R 1-x A x MnO 3 perovskite manganites, where R and A are some rare<br />

earth and alkaline earth elements respectively and 0.2 < x < 0.5, display the<br />

unusual property of being paramagnetic insulators at high temperatures and<br />

ferromagnetic metals at low temperatures [1-4]. Perovskite is the name of the<br />

structure type, Figure 1-1, containing corner sharing MnO 6 octahedra. Both<br />

end members of La 1-x A x MnO 3 are antiferromagnetic insulators [5], but become


4 Chapter 1<br />

ferromagnetic metals upon doping. The theory of double exchange [6-8],<br />

described in section 1.3, has been developed in order to explain this<br />

phenomenon and correctly predicts x = 1/3 to be optimal doping [9]. Recent<br />

calculations show that a second mechanism such as a Jahn-Teller distortion<br />

may be required to explain the magnetoresistance within the double exchange<br />

model [10-12].<br />

Until recently, much of the experimental work on the manganites has<br />

been motivated by their utility as a cathode materials in solid oxide fuel cells<br />

[13]. Thus many compounds of the type R 1-x A x MnO 3+δ have been studied in<br />

polycrystalline form [14-17]. Much has been learned about their defect<br />

chemistry and high temperature electronic and ionic conductivity. Most of<br />

these compounds are not metallic above room temperature but have<br />

electronic conductivity, presumably due to (small) polaron hopping,<br />

sufficient to make good electrodes.<br />

Interest in the perovskite manganites has expanded since their fabrication<br />

as epitaxial thin films [18, 19]. Some films have shown the insulator to<br />

ferromagnetic metal transition at lower temperatures with a large<br />

magnetoresistance near this transition [20, 21]. ΔR/R(H) of greater than 10 6 %<br />

has been reported for fields of several Tesla [22-25]. Since Giant Magneto<br />

Resistance (GMR) films have a ΔR/R(H) of typically 20% (which saturates in a<br />

few thousand Oersted), the manganite films have been proposed as possible<br />

replacements for GMR read heads in the magnetic recording industry.<br />

However, since magnetic recording devices work at room temperature with<br />

low magnetic fields, the temperature range and field sensitivity of the<br />

manganites in their present state do not make them competitive with GMR<br />

materials. Nevertheless, the rather imprecise term "Colossal Magneto<br />

Resistance" (CMR) has been coined for this phenomenon.* However, since it<br />

* It should be noted however, that such a large magnetoresistance is not


Introduction 5<br />

has been widely adopted it will be employed it here where CMR is defined as<br />

ΔR/R(H) > 10. CMR materials often refers to all manganite perovskites.<br />

Although the films are quite stable and the measurements reproducible<br />

even after several months, it is clear that growth and annealing conditions<br />

greatly influence the properties of the manganite films [27]. Furthermore, the<br />

electrical and magnetic properties of the CMR films are often very different<br />

than those of the materials produced by bulk ceramic techniques or single<br />

crystals with the same nominal composition. Thus, in order to understand<br />

these materials, one should distinguish between the properties intrinsic to<br />

perfect crystalline R 1-x A x MnO 3 and those caused by microstructure, strain,<br />

disorder and/or compositional variations.<br />

From the work described in chapter 4, it is concluded that the low<br />

temperature, CMR phenomenon is not intrinsic to the thermodynamically<br />

stable phases with composition La 0.67 Sr 0.33 MnO 3 or La 0.67 Ca 0.33 MnO 3 .<br />

In chapter 5 the effect of the rare earth magnetism is shown for the case<br />

R = Gd in Gd 0.67 Ca 0.33 MnO 3 . The possibility of structural distortions at T C are<br />

considered for this compound.<br />

1.3 Double Exchange<br />

The theory of double exchange is concerned with the exchange process<br />

involving d-band carriers in a mixed valent oxides. First postulated by Zener<br />

3+ 2+<br />

3+ 4+<br />

[6] to explain the properties of ( La1− A )( Mn1− Mn )O<br />

x x x x 3 [1, 2, 4], the theory of<br />

double exchange was formulated by Anderson and Hasegawa [7] and<br />

DeGennes [8]. The compounds at the two ends of the series are<br />

unique to the manganates. Doped EuO and EuS show magnetoresistances<br />

of 10 4 %, using the above definition, and therefore can be considered a<br />

CMR material. Furthermore, it has been shown that in some Chevrel<br />

phase compounds [26], a magnetic field makes the material<br />

superconducting - which would make them Òsuper-magnetoresistanceÓ<br />

(SMR) materials.


6 Chapter 1<br />

e g<br />

t 2g<br />

e g<br />

t 2g<br />

CaMnO 3 T


Introduction 7<br />

For x = 0, the situation is slightly more complex. Most reports claim that<br />

stoichiometric LaMnO 3 is an antiferromagnet insulator [29]. It is apparently<br />

difficult to prepare stoichiometric LaMnO 3 which likes to lose oxygen or be<br />

rich in lanthanum. Off-stoichiometric LaMnO 3 will contain mixed-valent<br />

manganese and could then be metallic and ferromagnetic. Stoichiometric<br />

LaMnO 3 contains entirely Mn 3+ which has 4 d electrons. The first 3 fill the<br />

spin-up t 2g band, as in CaMnO 3 , while the remaining electron half-fills the<br />

spin-up e g band. The e g band is apparently further split, resulting in an<br />

insulator. There are several ways the band could be split, any or all of which<br />

may be the cause of the insulating behavior. First of all, a half-filled band is<br />

susceptible to splitting due to the Mott correlation effect Ð producing a Mott<br />

insulator. Secondly, the structure is not entirely cubic particularly for the end<br />

members. Such a distortion raises the degeneracy of the t 2g and e g orbitals.<br />

This is known as a Jahn-Teller splitting. Finally, the unit cell relevant to the<br />

electronic structure may be doubled, which will split the e g band in half. The<br />

magnetic structure, by virtue of the antiferromagnetism, has a doubled cell,<br />

which may affect the electronic structure.<br />

At finite values of x there will be x holes (or 1 - x electrons) in the spin-up<br />

e g band. These holes should be free to move and provide a large conductivity.<br />

If, however, the intra-atomic exchange, which holds the spins of all the d<br />

electrons on a given ion parallel, is stronger than the Òhopping integral,Ó<br />

then the hopping can only take place between pairs of ions on which the t 2g<br />

spins (Mn 4+ core) are parallel. Otherwise the two sites have different energies.<br />

The difference in energy is proportional to -cos(θ/2), where θ is the angle<br />

between the neighboring core spins. Since free carriers gain kinetic energy by<br />

being itinerant, this provides a type of exchange mechanism which holds the<br />

two core spins parallel. Conversely, the more parallel the core spins are<br />

aligned, the easier it is for the carriers to become itinerant. Since a magnetic<br />

field has a large effect of aligning ferromagnetically coupled magnetic spins


8 Chapter 1<br />

near T C (magnetic susceptibility becomes large), the application of a magnetic<br />

field should increase the conductivity near T C . This gives a simple qualitative<br />

explanation for the large negative magnetoresistance observed near T C .<br />

It has been suggested the double exchange mechanism alone cannot<br />

provide such a large effect on the resistance [10]. It is proposed, that the<br />

electron-phonon coupling which localizes the conduction electrons as<br />

polarons at T > T C , augments the double exchange mechanism to provide the<br />

observed effects [11, 12]. This conclusion is not universally accepted [9, 30-32].<br />

The polaronic mechanism alone may account for similarly large<br />

magnetoresistance in ferromagnetic semiconductors [30, 33, 34]. The stable<br />

state of a electron donor in a ferromagnetic semiconductor can abruptly shift<br />

from being a shallow to a deep donor as the temperature is raised toward T C .<br />

The increasing spin disorder destabilizes the large-radius donor, which<br />

collapses into a well localized small-polaronic donor. The electron-lattice<br />

interaction plays a pivotal role in this phenomenon. The magneto-resistance<br />

arises because the temperature of the donor-state collapse and the<br />

accompanying metal-insulator transition are increased by the application of a<br />

magnetic field. Other explanations for magnetoresistance in ferromagnetic<br />

materials are discussed in chapter 6.<br />

When superexchange is of comparable magnitude to the double exchange,<br />

a canted antiferromagnetic ground state is expected. This is because<br />

superexchange favors an antiferromagnetic ground state with energy<br />

proportional to cos(θ) while the double exchange is proportional to -cos(θ/2).<br />

The minimum of these two energies is in general some θ ≠ 0 [8].


2. Materials Synthesis and Characterization<br />

In order to find or understand new physical phenomena, samples for<br />

measurement need to be made. The advantage of a materials physicist who is<br />

also a materials chemist is that he/she has more control over the material.<br />

There are many synthetic details which may effect the properties. Also, being<br />

able to make oneÕs samples makes it much easier to chose what materials to<br />

study and get the research started rapidly. A materials chemist who is also a<br />

materials physicist understands what properties of the material is of interest<br />

and what measurements are simple enough to characterize the samples so<br />

that chemical improvements can be made efficiently and effectively.<br />

2.1 Sample Preparation<br />

Sample preparation is sometimes viewed as a black art, or as simple as<br />

making breakfast. Indeed some materials require an immense amount of<br />

time and equipment. However, both of these requirements can be limited if<br />

the sample requirements are not too stringent. Much time and effort can be<br />

saved if the type of sample made just exceeds the sample requirements. Bulk<br />

polycrystalline samples are usually quite easy to make while growing crystals<br />

is more risky and time consuming [35]. When looking for isotropic<br />

properties, polycrystalline samples usually suffice.<br />

2. 1. 1 Bulk Polycrystalline Samples<br />

The most widely used method for preparing polycrystalline oxides is the<br />

direct reaction, in the solid state, of a mixture of solid starting materials.<br />

Powder solids are formed which can then be pressed and sintered to form<br />

dense polycrystalline pellets. Even though the desired phase is<br />

thermodynamically favored, solids do not usually react together at ambient<br />

temperature over laboratory time scales and it is necessary to heat the<br />

reactants at high temperatures to overcome the kinetic barriers. For such<br />

9


10 Chapter 2<br />

reactions, the rate limiting step is usually the solid state diffusion of the<br />

cations across the interface between the starting materials. In order to supply<br />

sufficient thermal energy to enable the ions to jump out of their normal<br />

lattice sites and diffuse through the crystal, high temperatures usually greater<br />

than 1000°C are required. Even at these temperatures, diffusion lengths are<br />

usually quite short. To facilitate this process, the starting materials are<br />

usually ground to a fine powder, which both decreases the length the ions<br />

must travel and increases the surface area for reaction. The powders are often<br />

pressed into a pellet before heating to increase the contact between particles.<br />

Reaction times are usually several days and it is best to repeat the process to<br />

insure homogeneous samples.<br />

Starting materials are usually single cation oxides, carbonates, nitrates or<br />

hydroxides Ñ materials which decompose to form oxides when heated.<br />

Carbonates are popular for the alkali and alkaline earth elements because they<br />

are not hydroscopic and therefore can be weighed accurately in air. On the<br />

initial heating or calcination of carbonate containing mixtures, carbon dioxide<br />

is produced and escapes from the solid. This prevents good sintering of the<br />

material into a dense ceramic, requiring an additional heating.<br />

At such high temperatures, the reactivity of the crucible material must be<br />

considered. Common crucible materials for high temperature reactions are<br />

alumina, zirconia, magnesium oxide and platinum. These materials may<br />

contain other impurities to help in their processing. So, contamination of the<br />

desired product by the crucible may not necessarily be by Al, Zr, or Mg.<br />

2. 1. 2 Single Crystals<br />

Many measurements of materials properties are easier to interpret if single<br />

crystal samples are used. Transport measurements, which require a<br />

contiguous transport pathway across the sample, can therefore be greatly<br />

influenced by the presence of grain boundaries, interfacial impurity phases<br />

and voids which obstruct or alter the paths. For instance, an impurity phase


Materials Synthesis and Characterization 11<br />

(of too little volume fraction to detect by many of the characterization<br />

techniques described in section 2.2) located in the grain boundaries may<br />

unknowingly dominate the resistivity.<br />

Furthermore, many properties of interest are tensorial in nature and<br />

therefore have some degree of anisotropy. The properties of polycrystalline<br />

samples are artificially isotropic due to the random orientation of the<br />

crystallites. This may give a easy way to measure the average property of a<br />

material but may be misleading in two ways. First, the physics of the material<br />

may depend largely on the anisotropy. For example, some materials such as<br />

graphite are metallic in one direction and insulating in another. Secondly,<br />

for practical applications most anisotropic materials experience some kind of<br />

orientation during the materials processing e.g. rolling, extruding, thin film<br />

deposition. Thus, the properties of the final product may strongly depend on<br />

the processing, because of the anisotropy.<br />

2. 1. 2. 1 Flux Growth<br />

The use of a homogeneous, amorphous solution may greatly facilitate<br />

formation of the crystalline product, since convection not diffusion will<br />

transport the ions, and the product will form at much lower temperatures<br />

than by solid state reaction. Flux growth can also sometimes yield metastable<br />

phases which are difficult or impossible to prepare by other means. The<br />

solvent or flux can be any material which dissolves and precipitates the<br />

desired material (solute). At the beginning of the crystal growth, the solute is<br />

entirely dissolved in the solvent. The solubility of the solute in the solvent is<br />

then decreased (usually by decreasing the temperature), causing crystals to<br />

nucleate and grow. Knowledge of the solute-solvent phase diagram will help<br />

determine appropriate concentrations and temperature ranges for crystal<br />

growth. The cooling rate and temperature gradient regulate the number and<br />

size of the crystals grown. If a gaseous transporting agent is used (vapor phase<br />

transport), the crystals will not have to be physically separated from the flux.


12 Chapter 2<br />

Otherwise one has to seriously consider how to remove the solvent after the<br />

crystals are grown. One should also be aware of possible contamination in the<br />

crystal by the flux or crucible material.<br />

For instance, crystals of SrRuO 3 and CaRuO 3 can be grown from SrCl 2 or<br />

CaCl 2 molten salts respectively [36]. The magnetic properties of such SrRuO 3<br />

crystals is described in Appendix A. During some of the growths, crystals with<br />

very different morphologies were found. Microprobe analysis showed<br />

significant platinum contamination in these crystals, presumably from the<br />

crucible. The Pt containing crystals from the SrRuO 3 growth were pyramidal<br />

and paramagnetic not ferromagnetic.<br />

The Pt containing crystals from the CaRuO 3 (denoted ÒCaRu/PtO 3 Ó) had<br />

cubooctahedral or rhombohedral morphologies. X-ray diffraction of selected<br />

single CaRu/PtO 3 crystals had a perovskite unit cell. X-ray diffraction of<br />

powdered crystals showed both perovskite and possibly Ca 4 (Ru/Pt)O 6 [37].<br />

This different phase was discovered also from an attempted crystal growth<br />

with CaCl 2 in a Pt crucible.<br />

2. 1. 2. 2 Float Zone<br />

Congruently melting compounds (materials which melt before<br />

decomposing) are ideal for float zone crystal growth. In this method no flux<br />

or crucible is used, preventing possible contamination problems. A<br />

polycrystalline source rod is made stoichiometricly. A molten zone then<br />

slowly moves down the rod. The molten zone is small enough that the<br />

surface tension of the liquid keeps it suspended between the two solid<br />

sections. The composition can be adjusted for growing non-congruently<br />

melting compounds. In this case, the traveling molten zone would have a<br />

composition different (perhaps containing a flux) from the desired material.<br />

The system used for the work reported here, uses a CO 2 laser focused to about<br />

1mm 2 to heat the molten zone.


Materials Synthesis and Characterization 13<br />

2. 1. 2. 3 Thin Films<br />

Related to vapor growth, is the growth of thin films. Thin films are easy<br />

to manipulate and therefore can be much more useful to electronics<br />

technologies than single crystals. The transport of the material to the<br />

substrate usually takes place in the gas phase or a vacuum. Once the atoms or<br />

molecules hit the substrate they diffuse only until they lose energy and are<br />

incorporated into the solid film. These atoms have a relatively short time to<br />

grow crystals and cannot usually return to the vapor phase. This is in<br />

contrast to standard crystal growth which relies on the solid-fluid equilibrium<br />

to grow single crystals. Nevertheless, single crystal films can be grown when<br />

the substrate is itself a single crystal.<br />

The various thin film deposition techniques differ primarily by the way<br />

the material is transported and how it gets into the vapor phase. The most<br />

common methods used in research transport the material in a vacuum or<br />

low pressure gas. The atoms, ions, or small inorganic molecules are ejected<br />

into the gas phase by evaporation, sputtering or laser ablation of a target.<br />

Chemical vapor deposition uses volatile precursors which can easily be<br />

transported in the gas phase to the hot substrate where they decompose to<br />

make the film. Many of the films used in this work were produced using<br />

solid-source MOCVD. In this case, a solid organometallic compound is<br />

evaporated, transported in the vapor phase to the substrate where it<br />

decomposes to form the film. Oxide films can also be produced by first spin-<br />

coating a sol-gel mixture of the desired metal stoichiometry. During heating,<br />

the sol-gel decomposes and crystallizes to form the film.<br />

2. 1. 3 Reactive Samples<br />

Many compounds react with water, oxygen or nitrogen present in air.<br />

Materials such as the subnitrides of barium [38, 39] or alkali intercalated C 60<br />

[40] decompose rapidly in air. Thus the synthesis and analysis of these<br />

materials is much more complicated since it must be done in an inert


14 Chapter 2<br />

atmosphere or vacuum. A glove box or Schlenck apparatus allows the<br />

manipulation of samples in an inert (usually argon gas) atmosphere.<br />

Samples can then be sealed in glass for analysis.<br />

Even if the sample is reasonably stable in air at room temperature, almost<br />

all non-oxides will react with oxygen if heated to high temperatures. Thus<br />

compounds such as sulfides [41, 42] or nitrides [43] must be sealed in glass<br />

before heating to solid-state diffusion temperatures.<br />

2.2 Characterization<br />

2. 2. 1 Elemental Analysis<br />

There are various ways to confirm or measure the elemental composition<br />

of a sample. Most of these methods utilize properties of the core electrons or<br />

nuclei and tend to be insensitive to the light elements. For elements heavier<br />

than neon, there exists several accurate and common procedures: The<br />

electron microprobe detects characteristic X-rays produced after excitation by<br />

an electron beam in an electron microscope. The characteristic light emission<br />

or absorption of gaseous elemental species can be used for elemental analysis<br />

Ð one common method uses an inductively coupled plasma (ICP).<br />

Rutherford back scattering (RBS) detects the weight and depth of atoms by<br />

elastic scattering of alpha particles off their nuclei. X-ray diffraction can in<br />

principle be used for elemental analysis since it gives essentially an electron<br />

density map. Wet chemical methods such as iodometric titration can be used<br />

to measure the oxidation state of individual elemental species, allowing the<br />

inference of the oxygen stoichiometry.<br />

2. 2. 2 Structural Analysis<br />

The physical properties of a material may depend as much on the<br />

structure as the elemental composition: graphite is quite different from<br />

diamond. The regular periodic structure can be determined from the elastic<br />

scattering of neutrons or X-rays. Wavelengths approximately equal to the


Materials Synthesis and Characterization 15<br />

interatomic spacings (measured in ) are needed. Synchrotron and neutron<br />

sources usually have tunable wavelengths as well as high intensity. For<br />

many purposes, however, laboratory scale X-ray diffraction is often adequate.<br />

2. 2. 2. 1 Neutron diffraction<br />

Neutrons, having a magnetic moment, are sensitive to the magnetic<br />

structure as well as the atomic structure. For example, an antiferromagnet<br />

which has a larger magnetic unit cell than the atomic cell, will cause extra<br />

neutron diffraction peaks not seen in X-ray diffraction. Polarized neutrons<br />

are also sensitive to the orientation of these magnetic moments. Since<br />

magnetic measurements only give the net moment, which is often zero for<br />

an antiferromagnet, neutron diffraction is far superior in determining<br />

magnetic structures. Inelastic neutron diffraction can provide further<br />

information concerning the structure dynamics. For instance phonon and<br />

magnon (spin wave) dispersions can be measured.<br />

2.2.2.2 X-ray diffraction<br />

X-ray diffraction (XRD) can be used both to quickly determine which<br />

phases are present in an unknown sample and to perform a detailed<br />

structural investigation. The difference lies more in the sample preparation<br />

and data analysis than the measurement apparatus. Although synchrotron<br />

source x-ray diffraction experiments are quicker, much of what is desired can<br />

be learned from a standard laboratory experiment.<br />

X-rays scatter off the periodic arrays of atoms in a crystal lattice. The<br />

scattered x-rays produce a pattern unique for a particular substance. This<br />

pattern can be used for either phase identification or it can be analyzed to<br />

determine the position of the atoms in the cell. The orientation of the crystal<br />

with respect to the incoming x-rays determines the orientation of the<br />

diffracted x-rays. So, either a single crystal must be precisely oriented to detect


16 Chapter 2<br />

a particular diffraction peak, or many crystallites randomly oriented can be<br />

used to get an orientation independent response.<br />

2.2.2.2.1 Powder X-ray diffraction<br />

In powder x-ray diffraction, the sample consists of many small crystallites<br />

which are assumed to be randomly oriented. This produces rings of diffracted<br />

x-rays, defined by the angle between the incident and diffracted beams, 2θ.<br />

The diffraction condition is determined by the atomic unit cell. In this way,<br />

the structure type of a material and unit cell size can easily be determined. If<br />

more than one type of material is present, the diffraction pattern will be a<br />

superposition of each of the components. If only phase identification is<br />

desired, this often provides enough information. Reference [40] gives a good<br />

example of how this is done.<br />

The intensities of each diffraction line is determined by the atomic<br />

constituents, and their placement in the unit cell. In principle, both of these<br />

can be determined from the intensities. In practice, this is difficult to achieve.<br />

A primary concern is texturing, or preferred orientation of the crystallites. If<br />

the crystallites are plate or needle like, they will tend to lay in the sample<br />

holder in a non-random orientation. This will cause a variation in the<br />

intensities. Also, individual diffraction peaks often overlap one another in a<br />

powder pattern. Since the determination of the structure depends on how<br />

much intensity is associated with each peak, having overlapping peaks<br />

complicates the solution process. It is often easiest to do structural analyses<br />

on single crystal samples.<br />

Powder diffraction sample holders are usually for flat, planar samples. If<br />

the incident angle of the incoming x-rays with respect to the sample plane is<br />

equal that of the diffracted beam being detected, the x-ray beam is partially<br />

focused (Bragg-Botano parafocusing) to give a narrower diffraction peak. For<br />

this reason, the sample plane is usually rotated (by an angle ω = θ) as the


Materials Synthesis and Characterization 17<br />

detector rotates (by angle 2θ). Plate-like crystallites can be highly oriented in<br />

such a sample holder. The Guinier method utilizes a different focusing<br />

technique for samples placed in a capillary tube. This is ideal for air sensitive<br />

samples which can be easily sealed in a tube. Needle like crystallites are easily<br />

oriented when placed in a tube.<br />

2.2.2.2.2 Single crystal and films<br />

Single crystals diffract an x-ray beam to produce spots as opposed to line. A<br />

particular spot appears only when the crystal is in a particular orientation.<br />

For this reason, three additional angles are used to orient the crystal. If the<br />

crystal is very large, it will absorb much of the x-ray intensity. Small crystals,<br />

about 0.2mm diameter, are usually used for structure determination since<br />

they absorb little and have a constant volume of sample in the x-ray beam at<br />

all time. The crystal can be rotated and the intensity of each diffraction spot<br />

measured. This usually provides enough data so that the independent data to<br />

free parameters ratio is about 10. Examples of structure determinations using<br />

single crystal x-ray diffraction can be found in [38, 39, 41, 42, 44-48]. If there are<br />

many smaller crystals in the x-ray beam, they can be effectively ignored by<br />

measuring x-ray intensity only at positions where a diffraction peak from the<br />

larger crystal is expected. This technique is used in [38, 39, 45, 46].<br />

Single crystal films also produce diffraction spots as opposed to lines.<br />

However, the intensities are small compared to those from the substrate. It is<br />

usually only possible (and only of interest) to determine filmÕs unit cell size,<br />

orientation and texturing. The difference in cell size of the film compared to<br />

a bulk sample gives an indication of the strain on the film.<br />

2. 2. 2. 3 X-ray Absorption Fine Structure<br />

X-ray Absorption Fine Structure (XAFS) gives essentially the pair<br />

distribution function of atoms around a particular element in the sample.


18 Chapter 2<br />

Thus XAFS probes the local structure, as compared to X-ray diffraction which<br />

gives average structure over hundreds of angstroms.<br />

X-ray absorption spectra in this study were collected by Corwin Booth and<br />

Bud Bridges from U. C. Santa Cruz in transmission mode on beam line 4-3 at<br />

the Stanford Synchrotron Radiation Laboratory (SSRL) using powder samples<br />

(grain size less than 30μm). The sample temperature was regulated using an<br />

Oxford helium cryostat system within 0.1 K (absolute temperature may be as<br />

much as 2 K warmer). Data for this experiment were collected above and<br />

below T C , at T=69 K and T=40 K. Data reduction and analysis followed<br />

standard procedures reported previously [49].


3. Electronic and Magnetic Measurements<br />

3.1 Transport Properties<br />

Transport properties: Resistivity, dielectric constant, thermalconductivity,<br />

thermopower, Hall effect, etc. are often a prime concern when engineering a<br />

new material for a particular use. Even if the purpose of the material is not<br />

related to transport, the material may still be required to have certain<br />

transport characteristics. For example, the liquid crystal in a liquid crystal<br />

display should have a low electrical conductivity to minimize resistive losses.<br />

3. 1. 1. 1 Ohm’s Law<br />

Most materials, whether metals, semiconductors or insulators, obey<br />

OhmÕs law to a good extent: the current I flowing in a wire is proportional to<br />

the potential drop V along the wire V = IR (linear response). R is the<br />

resistance of the wire and it depends on the size and shape of the wire. One<br />

generally prefers to use intensive quantities to characterize a material. Thus<br />

OhmÕs law can be used to define the resistivity ρ which is defined to be the<br />

proportionality constant between the electric field E and the current density j<br />

that it induces: E = ρj. Since E and j are vectors, ρ is a second order tensor.<br />

Any deviations of OhmÕs law can be easily described by adding terms with<br />

higher powers of j. The conductivity σ is the inverse of the resistivity (σ = ρ -1 )<br />

such that j = σE. In an isotropic or cubic substance the resistivity tensor ρ has<br />

off diagonal elements equal to zero and three equal diagonal elements ρ (a<br />

scalar), then the conductivity tensor σ has the same form with diagonal<br />

elements σ where σ = 1/ρ. For an orthorhombic substance (for which ρ and σ<br />

have no nonzero off diagonal terms in the appropriate coordinate system),<br />

the conductivity along a principle direction is equal to the inverse of the<br />

resistivity along the same principle direction. Because of these simple<br />

19


20 Chapter 3<br />

relationships resistivity and conductivity are commonly described as if they<br />

were scalars. Some other properties of the conductivity and resistivity tensors<br />

are described in section 6.2.1.<br />

3.1.1.2 Magnetoresistance<br />

Magnetoresistance is simply the change in resistivity as the magnetic field<br />

is applied. For nonmagnetic metals, the magnetoresistance ratio ΔR/R is only<br />

a few percent in large ~1Tesla fields. For symmetry reasons discussed in<br />

section 6.2.1, the magnetoresistance is proportional to H 2 for small H in these<br />

metals. Since this magnetoresistance arises from the complicated orbits of the<br />

electrons on the Fermi surface, the magnetoresistance has large<br />

crystallographic anisotropy.<br />

Even in an isotropic conductor, there exists two possible configurations for<br />

magnetoresistance. When the current is parallel to the applied magnetic field<br />

the longitudinal magnetoresistance is measured. If the magnetic field is<br />

perpendicular to the current path, then the transverse magnetoresistance is<br />

being measured.<br />

If the material itself is not isotropic, different crystallographic orientations<br />

will have distinct longitudinal and transverse magnetoresistances. For<br />

example, a cubic material grown as a thin film may have growth induced<br />

crystallographic anisotropy. Since for practical purposes, the current is<br />

usually constrained to the plane of the film, there are three easily reportable<br />

magnetoresistances: longitudinal and transverse magnetoresistance with<br />

both the current and magnetic in the plane of the film, and transverse<br />

magnetoresistance with the magnetic field perpendicular to the film. The<br />

two different magnetic field orientations will also affect the demagnetization<br />

field and domain size and motion which must be taken into account.<br />

Magnetoresistance that is actually proportional to the magnetization<br />

rather than the magnetic field is called anisotropic magnetoresistance (AMR).<br />

In low fields, the magnitude of M does not change, only the orientation. If M


Electronic and Magnetic Measurements 21<br />

is aligned parallel (or antiparallel) to the current, the resistance is different<br />

than if M were perpendicular to the current. This difference is the anisotropic<br />

magnetoresistance. AMR in La 0.67 Ca 0.33 MnO 3 will be discussed further in<br />

sections 6.1.<br />

3. 1. 1. 3 Drift velocity, mobility, relaxation time and mean free path<br />

It is often useful to think of transport in terms of particles moving with an<br />

average velocity in a field because of collisions with the surroundings. Some<br />

of the basic relationships are given below for electrical conductivity, and can<br />

be easily generalized for other transport phenomena. The electrical current j<br />

(charge per cross-sectional area per time) is related to the number density of<br />

carriers n, the charge of each carrier -e (for electrons), and the average drift<br />

velocity v.<br />

j = -nev.<br />

The drift mobility μ (a tensor like the conductivity) is defined by the<br />

relationship between the drift velocity and the electric field E.<br />

v = μE<br />

Thus the conductivity is proportional to the number of carriers and their<br />

mobility<br />

σ = -neμ<br />

The relaxation time can also be used to relate the drift velocity and the<br />

electric field. The electric force accelerates the carrier, of mass m, for an<br />

average time τ between collisions (the relaxation time) to provide the drift<br />

velocity.<br />

v = E<br />

−eτ<br />

τ<br />

so σ =<br />

m<br />

ne<br />

2<br />

m<br />

The mean free path l, the average distance the carrier travels between<br />

collisions, is a useful quantity. If the mean free path is smaller than the<br />

interatomic spacing a, then clearly a model based on atomic collisions is


22 Chapter 3<br />

inadequate. l ≈ a is known as the Ioffe-Regel limit. To calculate l, the average<br />

speed (not velocity) must be known. Unfortunately, this is the Fermi velocity<br />

which depends strongly on the band structure. For an electron gas of density<br />

n with a spherical Fermi surface, the Fermi velocity v F is<br />

v<br />

.<br />

F ≈ 309 3<br />

m n<br />

h<br />

h σ<br />

so that l ≈ 309 . 2 2 .<br />

3 e n<br />

3. 1. 1. 4 Hall effects<br />

The Hall effect is the transverse electric field that is produced by the<br />

presence of a magnetic field perpendicular to the flow of charged carriers.<br />

This can be described by antisymmetric off diagonal terms in the resistivity<br />

tensor which are a function of the magnetic field. The Hall effect is normally<br />

linear with respect to an applied field. For an isotropic substance E y = j x R H H,<br />

where H is the magnetic field, j x is the current along the x direction and E y is<br />

the transverse electric field. R H is known as the Hall coefficient. An internal<br />

magnetic field due to a nonzero magnetization of the material may cause a<br />

Hall effect. This anomalous Hall effect is proportional to the internal<br />

magnetization M, E y = j x R A M. The angle between E and j caused by the Hall<br />

effect is known as the Hall angle.<br />

For simple metals and semiconductors, the Hall effect (combined with<br />

other data) can reveal the sign and density of the charge carriers. The Hall<br />

coefficient for an electron gas is -1/nec. If there is more than one type of<br />

carrier, the Hall effect is the weighted average of the Hall effect of the<br />

individual carriers and becomes nonlinear in H. This nonlinearity can be<br />

used to determine the sign and density of the individual carriers Ð assuming<br />

field independent mobilities and Hall coefficients for each carrier type [50].<br />

3. 1. 2 Measurement<br />

The primary concern for reliable transport measurements of new<br />

materials is the identification and elimination of systematic errors. Thermal<br />

and electrical noise lower the precision but usually not to an important extent


Electronic and Magnetic Measurements 23<br />

when characterizing new materials. The danger is finding a new effect or<br />

result that is reproducible but none the less spurious. Since measurements of<br />

new materials have usually not been previously performed, it is often<br />

impossible to compare with previous results. Some considerations and<br />

solutions are outlined below.<br />

3.1.2.1 Linearity<br />

Most transport measurements assume some sort of linear response.<br />

Conductivity measurements assume I is linearly proportional to V.<br />

Experimentally this is never exactly true. A test for linearity should always be<br />

made before measuring a new sample. The resistance over several orders of<br />

magnitude should be constant within a few percent. Even if I vs. V is linear,<br />

thermal or contact voltages usually add a nonzero offset voltage V 0 so that V<br />

= IR + V 0 . Using wires from the same spool, pure copper contacts and<br />

shielding may help but will not totally eliminate V 0 . This offset voltage is<br />

easily subtracted. The easiest method is by measuring the voltage at I and -I.<br />

The resistance is then R(I) = (V(I) - V(-I))/2. The best method is to take a full I<br />

vs. V curve and measure the slope. This may take more time and therefore<br />

introduce other errors such as those caused by temperature drift. The current<br />

range used for measurement should be well within the Ohmic regime. A<br />

current too small may have V 0 comparable to V; it is best not to rely heavily<br />

on the subtraction of V 0 . Large currents may produce I 2 R heating of the<br />

sample, in which case the measured resistance should look like it has an<br />

additive term proportional to I 2 . AC measurements are ideal for extracting<br />

only the linear response; in this case, the frequency dependence should also<br />

be checked, and compared to the DC value.<br />

Thermopower S measurements similarly assume the linearity of the<br />

voltage with the temperature difference: V = SΔT. These measurements also,<br />

however, have an offset voltage V 0 that should be subtracted.


24 Chapter 3<br />

Hall effect measurements have several assumptions which need to be<br />

considered. First, since it is basically a resistance measurement, the I vs. V<br />

curve should be checked and V 0 subtracted as above. Second, it is often<br />

assumed that the Hall voltage is linear in the applied magnetic field<br />

E y = j x R H H. This is often not the case, as described above (section 3.1.1.4) for<br />

magnetic materials or semiconductors with more than one type of carrier.<br />

Finally, the effect of the magnetoresistance needs to be subtracted.<br />

Traditionally, and theoretically this can be done by balancing the Hall voltage<br />

so that it reads zero at H = 0, and then the magnetoresistance should also be<br />

subtracted for H ≠ 0. In practice, this does not work. There is usually some<br />

magnetoresistive component in the measurement perhaps due to some<br />

anisotropy or inhomogeneity in the material. Since the magnetoresistance is<br />

an even function of H while the Hall effect is an odd function of H, I have<br />

found it easiest to examine the Hall effect by plotting V y (H) - V y (-H) as a<br />

function of H. This give the antisymmetric portion of the Hall voltage<br />

(subtracting the symmetric magnetoresistive contribution) without assuming<br />

a Hall effect which is linear in the applied field H.<br />

3.1.2.2 Geometry<br />

In order to calculate the resistivity from the measured resistance, the<br />

geometric ratio relating the two must be known. Most analyses are based o n<br />

some temperature or field dependence, and therefore it is more important<br />

that this geometric factor does not change during the experiment than it is to<br />

know the value precisely. The geometrical factor can change if the sample<br />

has internal cracks or if the contacts are flowing or cracking.<br />

For resistivity measurements, the resistance of the entire circuit must be<br />

taken into account. A tiny metallic sample may have a resistance much<br />

smaller than that of the contacts or even the wires leading to the voltmeter.<br />

For this reason, four probe measurements are used. A known current is<br />

applied through the sample from two of the contacts. The voltage across a


Electronic and Magnetic Measurements 25<br />

portion of the sample is sensed with two other contacts. The voltmeter has a<br />

high input impedance so that very little current is drawn from the voltage<br />

contacts. If there is almost no current flowing through the voltmeter circuit,<br />

then there is almost no voltage drop across the contacts or contact wires. In<br />

this way, the voltage measured is the voltage across the sample. Since the<br />

voltage contacts are separated from the current contacts, one must be cautious<br />

of the tacit assumption that the current flows uniformly through the entire<br />

sample. If the current avoids the voltage contacts, the results will be<br />

spurious. For high resistance samples, two probe measurements, where the<br />

voltage contacts are the same as the current contacts, can be used. The<br />

resistance due to the contacts and grain boundaries can be determined with<br />

AC impedance spectroscopy. Extremely high resistance measurements can be<br />

done with an electrometer where a constant voltage is applied and the<br />

electrometer measures the tiny trickle of current that passes through.<br />

The simplest geometry is the bar sample. The resistivity of a bar is the<br />

resistance times the cross-sectional area divided by the length between the<br />

voltage contacts. Samples can either be shaped into bars or rods, or films can<br />

be patterned. Patterned films can be ideal for transport measurements since<br />

the geometric ratio is well known, and the effects of the contacts can be<br />

minimized by patterning small contact wires in the film. Unfortunately the<br />

patterning process can damage the films.<br />

For isotropic two-dimensional samples, the physical dimensions (other<br />

than the thickness) need not be measured. In some cases the geometric ratio<br />

can be calculated analytically using conformal mappings. The geometric ratio<br />

has been calculated for a sheet where four collinear and equally spaced<br />

contacts are in the center of a sheet [51]. The Van der Pauw [52, 53]<br />

configuration uses four contacts placed anywhere on the edge. By switching<br />

one of the current and voltage leads, the geometric ratio can be calculated.<br />

The Van der Pauw configuration can also be used for the Hall effect.


26 Chapter 3<br />

3.1.2.3 Contacts<br />

Bad contacts are often the cause of problems in a transport experiment.<br />

Making good contacts is the closest IÕve come to producing fine art in the<br />

laboratory. Contacts need to be strong and durable yet small. The thermal<br />

stresses of temperature cycling can weaken or break a contact. A weak contact<br />

is worse than a broken one because a weak contact always seems to work at<br />

room temperature, give unusual results at some other temperature and then<br />

return to normal at room temperature. A good contact is not delicate, it<br />

should be able to withstand some mechanical stress at room temperature.<br />

Contacts need to be small when it is assumed that they are point contacts<br />

or if the sample is physically small. If the size or shape of the contact affects<br />

the geometry of the measurement (section 3.1.2.2) any sintering, melting,<br />

cracking, etc. of the contact will affect the measurement. Unfortunately, the<br />

smaller the contact, the weaker it usually is.<br />

The contact wires can also cause stress in the contact since they also will<br />

expand and contract with temperature. Contact wires should have bends to<br />

minimize the strain on the contacts. Contact wires should, if possible, also be<br />

plastically deformed so that they will stay at the desired position without any<br />

mechanical force, before the contact material is applied.<br />

Indium metal makes good contacts for measurements at temperatures less<br />

than 400K. Indium is quite ductile and therefore can withstand considerable<br />

strain before breaking. Freshly cut indium can be pressed onto most flat<br />

sample surfaces. The contact wire can then be sandwiched between a second<br />

piece of indium. In this way, contacts a fraction of a millimeter can be made.<br />

For surfaces which are difficult to adhere to, an ultrasonic soldering iron<br />

usually helps. Shiny, ultrasonically vibrating, liquid indium will stick to<br />

almost anything and can be drawn out into long wires. Any yellow, oxidized<br />

surface of the molten indium will hinder the adhesion to a sample.


Electronic and Magnetic Measurements 27<br />

Silver paint contacts can be made quite small if the suspension has the<br />

appropriate viscosity. These contacts can be used to low temperatures but are<br />

often unreliable. I have had better results with the two component silver<br />

epoxy (epo-tek ¨ 417).<br />

Polyimide based silver epoxy (epo-tek ¨ P1011) can be used well above its<br />

recommended maximum temperature of 300°C. It requires only a low<br />

temperature curing and works almost up to the melting point of silver. This<br />

was used for measurements to 1200K. A platinum paste is available for high<br />

temperature contacts that requires a high temperature sintering.<br />

3.1.2.4 Reproducibility<br />

It is obviously important to only report reproducible data. Important data<br />

should always be reproduced and claims about general properties of a<br />

material should be reproduced on other samples or even types of samples. It<br />

is best to simply not collect irreproducible data. If there is anything unusual<br />

in the above checks, it is probably not worth the time collecting the data.<br />

Extensive automation of the data collection system may actually help<br />

increase the consistency of the measurement. If the temperature is changed<br />

in a uniform way, the data tends to be more consistent than if points were<br />

taken at random temperatures. This is probably due to temperature gradients<br />

in the sample or sample chamber, that depend on the heating or cooling rate.<br />

Measurements as a function of temperature should generally be measured<br />

while heating and cooling in case there is some temperature differential<br />

between the sample and the thermometer. Some contacts break easily when<br />

rapidly cooled, requiring the slow cooling of an automated system. With<br />

enough automation so that the experiment can be left unattended for several<br />

hours or days, extra data can be collected. This allows more frequent checks<br />

for linearity and reproducibility.


28 Chapter 3<br />

3.1.2.5 Apparatus<br />

Transport measurements for this work were made in DC allowing a few<br />

seconds for the current and voltages to stabilize. A constant current source<br />

(Keithley model 220 or 224) was used to supply the current and a separate<br />

voltmeter (Keithley model 181, 182, 196 or 197) used to measure the voltages.<br />

Current is not usually measured independently but can be with the addition<br />

of an electrometer into the circuit. A computer controlled the current source,<br />

read the voltages and temperatures and recorded the data. Typically, 5 pairs of<br />

R(I) = (V(I) - V(-I))/2 were measured at slightly different currents (e.g. 20%<br />

variation) to estimate the precision. For resistances greater than 10 6 Ω (often<br />

T < 100K) two point resistance measurements were made using an<br />

electrometer.<br />

Most of the low temperature and all the magnetoresistance and Hall effect<br />

measurements were made in the cryostat of the Quantum Design MPMS 2<br />

SQUID magnetometer described in section 3.2.1.1. The computer that controls<br />

the SQUID magnetometer can also operate subroutines that can control GPIB<br />

compatible devices. These EDC (External Device Control) subroutines can be<br />

quite general, but have a limited number of available commands. It is<br />

necessary to further process the data recorded from the EDC subroutine in<br />

order to use it in a graphing or spreadsheet program. The standard SQUID<br />

magnetometer software controls the temperature and magnetic field. The<br />

thermometer is not very near the sample, so reliable data must be taken after<br />

the temperature has stabilized. This usually takes about 5 minutes for<br />

routine measurements or 30 minutes if any temperature drift is to be<br />

avoided. Since the measurement can run for several days unattended, time is<br />

usually not much of a problem.<br />

A 10 wire transport probe prefabricated at Quantum Design was used for<br />

the measurements. A special tip was made out of 8 bore alumina rod to<br />

provide a surface perpendicular to the magnetic field. Films can then be


Electronic and Magnetic Measurements 29<br />

attached to this surface with double-sided sticky tape to measure the Hall<br />

effect and transverse magnetoresistance. A simple copper plate attached to<br />

the tip of the transport probe provides a surface parallel to the magnetic field<br />

for longitudinal and transverse magnetoresistance measurements. One<br />

limitation of the SQUID magnetometer is its relatively small sample chamber<br />

(8mm diameter).<br />

High temperature (300K < T < 1200K) measurements were performed in a<br />

furnace (in air) with a slowly varying temperature (≈ 1 K/min.). The<br />

temperature was recorded using a Type-K thermocouple positioned a few<br />

millimeters from the sample. The sample holder was made from an 8 bore<br />

alumina rod. The 2 thermocouple wires and 4 platinum wires which contact<br />

the sample were fed through the bores of the alumina rod. The wires are<br />

positioned to facilitate changing the sample and attaching new contacts. The<br />

furnace was programmed to heat and cool using a standard Eurotherm<br />

(model 818) power regulator and temperature controller. Data were<br />

continuously recorded by a computer using LabView software as the furnace<br />

heated and cooled. Lower temperature measurements can be similarly made<br />

by dipping the probe into a dewar of liquid nitrogen or liquid helium.<br />

3.1.3 Analysis<br />

Careful transport measurements can help identify transport mechanisms.<br />

Knowledge of transport mechanisms not only helps us to understand the<br />

microscopic nature of the material but also allows us to predict other<br />

properties using previous theoretical or empirical knowledge of other<br />

materials with the same transport mechanism. For example a few percent<br />

impurity due to processing may not affect the room temperature resistivity of<br />

a metal while greatly decreasing that of a band semiconductor.<br />

Some of the common transport mechanisms and their predicted transport<br />

properties are given below. Since there is not much overlap of mechanisms<br />

or properties between metals and insulators, it is convenient to categorize


30 Chapter 3<br />

materials as metals or insulators-semiconductors. Using the room<br />

temperature resistivity for this definition is attractive since metals are<br />

supposed to have low resistivity while insulators have high resistivity, but<br />

imprecise since a dirty semiconductor may have a lower resistivity than a<br />

poor metal at some high temperature. Even though theoretically metals<br />

have zero or low resistivity as the temperature approaches zero while<br />

insulators have infinite resistivity at T = 0, such a definition is impractical for<br />

materials which undergo a metal-insulator transition. Here, a metal is<br />

defined to be a material with a metallic slope in the resistance vs. temperature<br />

curve: dρ/dT > 0 while a insulator/semiconductor has dρ/dT < 0 This<br />

definition is consistent with most microscopic models of metals and<br />

insulators.<br />

3.1.3.1 Metals<br />

The resistivity of a metal is best described in terms of elastic and inelastic<br />

scattering mechanisms. Without scattering the itinerant electron<br />

wavefunctions would allow electron transport without loss or transfer of<br />

energy. In the relaxation time approximation, MatthiessenÕs rule applies,<br />

which states that the total resistivity is the sum of each of the individual<br />

resistivity contributions. The different scattering mechanisms usually<br />

depend on temperature with a particular power law. Thus the total resistivity<br />

is often a polynomial in T.<br />

3.1.3.1.1 Impurity scattering<br />

Impurities and defects disrupt the disrupt the Bloch wavefunctions and<br />

scatter electrons. Since this perturbation of the crystal lattice is unaffected by<br />

temperature, the resistivity is temperature independent. The impurities can<br />

be in the form of elemental substitutions on an atomic site, interstitial atoms<br />

or vacancies - anything that disrupts the periodic electronic structure. Since a


Electronic and Magnetic Measurements 31<br />

grain boundary is a plane of defects, grain boundary resistance is similarly<br />

temperature independent and often quite large.<br />

3.1.3.1.2 Electron-electron scattering<br />

The strong Coulomb interaction between electrons provides a mechanism<br />

for electron-electron scattering. The exclusion principle limits the scattering<br />

to electrons in partially occupied levels near the Fermi surface. For two<br />

electrons to scatter, both must be in this shell of partially occupied levels with<br />

width k B T about the Fermi level. This leads to a scattering rate and therefore<br />

resistivity proportional to (k B T) 2 .<br />

3.1.3.1.3 Electron-phonon scattering<br />

Other than impurities, lattice vibrations or phonons can disrupt the<br />

periodic structure of the lattice. This will scatter the electrons and lead to an<br />

intrinsic source of resistivity present even in a sample free of crystal<br />

imperfections. The most widely employed expression for this resistivity is<br />

the Bloch-GrŸneisen formula [54]: ρ =<br />

− − − ∫<br />

5<br />

5<br />

KT z dz<br />

6 z z<br />

Θ ( e 1)( 1 e )<br />

D<br />

Θ D<br />

T<br />

0<br />

where K is a<br />

constant characteristic of the metal, and Θ D is the Debye temperature (see<br />

section 3.3.2.2).<br />

At high temperatures T/Θ D > 0.5, this gives a resistivity proportional to T:<br />

2<br />

ρ ≈ KT/4ΘD . At low temperatures T/ΘD < 0.1 the resistivity follows the Bloch<br />

T 5 law: ρ ≈ 124.4 KT 5 6<br />

/ΘD .<br />

3.1.3.2 insulators/semiconductors<br />

A material is insulating or semiconducting when dρ/dT < 0. The<br />

mechanisms described below predict this behavior because the conduction<br />

must be thermally activated. The conducting species must overcome some<br />

energy barrier for conduction. Thermal energy supplies the energy to


32 Chapter 3<br />

overcome the barrier and therefore an increase in temperature increases the<br />

conductivity.<br />

3.1.3.2.1 Band insulators/semiconductors<br />

The electron transport in a semiconductor with an energy gap E g between<br />

the conduction and valence bands is limited by the number of excited carriers.<br />

In the intrinsic regime (low doping concentrations and high temperatures),<br />

the chemical potential lies in the middle of the gap so that the thermal energy<br />

required to excite a carrier is E g /2. The carrier concentration n thus increases<br />

exponentially with temperature n ≈ exp(E g /2k B T). The drift mobility μ = v/E<br />

defined by the drift velocity v in an electric field E in a semiconductor can be<br />

complicated and depends on the doping [55]. At low temperatures, impurity<br />

scattering is expected to dominate giving μ ∝ T 3/2 . Above about 100K lattice<br />

scattering will dominate with an approximate temperature dependence<br />

μ ∝ T -3/2 . Thus the conductivity of an intrinsic semiconductor is σ = neμ ≈<br />

T ν exp(E g /2k B T). For elemental semiconductors, the experimentally observed<br />

mobilities have an exponent ν = -1.5 or larger (more negative) [55]. The Hall<br />

effect is more complicated to determine a priori for an intrinsic<br />

semiconductor since the concentration of electron and hole carriers are the<br />

same.<br />

In the extrinsic regime (doped semiconductors at lower temperatures) the<br />

carrier concentration is determined by the thermal excitation of carriers from<br />

donor or acceptor impurities. The carrier concentration retains the form<br />

n ≈ exp(ΔE/k B T) where ΔE is determined by the difference in energy between<br />

the donor and band energies and the temperature regime of interest [56]. For<br />

example, ΔE is one half the band gap (E g /2) for an intrinsic semiconductor or<br />

ΔE = E g - E d ≈ E g for an extrinsic semiconductor with donor (or acceptor)<br />

energy levels E d from the band edge. Since usually one type of carrier (hole or<br />

electron) is dominant, the Hall effect R H = 1/nec is quite large since n is small,


Electronic and Magnetic Measurements 33<br />

varies exponentially with temperature and can be used to determine the sign<br />

of the carrier and the carrier concentration.<br />

3.1.3.2.2 Polarons<br />

A localized electron will always distort its surroundings relative to an<br />

unoccupied site simply because of the coulombic interaction of the electron<br />

and the surrounding atoms. The potential well produced by this distortion<br />

acts as a trapping center for the self-trapped carrier. The quasiparticle<br />

composed of a self-trapped electronic carrier taken together with the pattern<br />

of atomic displacements that produces the self-trapping became known as a<br />

polaron because self-trapping was first considered in ionic (polar) materials.<br />

The quasiparticle can move as a whole, the electron and the distortion<br />

moving together.<br />

The spatial extent of the self trapped states depends on the range of the<br />

interaction. Long-range electron-lattice interactions [57] produces large<br />

polarons with a finite radius. In such an instance, the self-trapped electronic<br />

carrier extends over multiple sites. The radius of the large polarons decreases<br />

continuously as the strength of the electron-lattice interaction is increased.<br />

The multi-site extension of the large polaron results in its itinerant motion.<br />

In contrast, small polarons are confined to a single lattice position due to<br />

strong short-range electron-lattice interaction. If the electron-lattice<br />

interaction is too weak, the carriers remain unbound. The extreme<br />

confinement of the small polaron typically leads to its moving by thermally<br />

assisted hopping.<br />

One of these two different types of polarons will be stable when both long-<br />

range and short-range electron-lattice interactions coexists. The presence of<br />

long-range interactions eases the requirements for forming a small-polaron,<br />

but only once the interactions are sufficiently strong will the polaron change<br />

from large to small.


34 Chapter 3<br />

The dichotomy between the two types of polarons seems unnecessary<br />

when considering their spatial extent, but the conductivity of the two types of<br />

polarons are vastly different. Large polarons move with significant mobilities<br />

μ > 1 cm 2 /V sec, that decrease with increasing temperature (metal-like). In<br />

contrast, small polarons move with very low mobilities, μ Θ D /2) the electrical<br />

conductivity is predicted [58] to be σ = 3ne 2 ωa 2 /2k B T -1 exp(-E A /k B T). Here n is<br />

the number density of charge carriers, a is the site-to-site hopping distance, ω<br />

is the longitudinal optical phonon frequency and e is the electronic charge.<br />

The Hall mobility may behave like the drift mobility with 1/3 of the<br />

activation energy, but in the adiabatic approximation the form is much more<br />

complicated [58].<br />

In the theory of small polaron hopping, it was found that the motion of<br />

the polaron is only weakly dependent on the magnetic order, and changes<br />

little as the temperature is raised above the Curie temperature [61, 62].<br />

Small polarons were first studied in the non-adiabatic regime, where the<br />

electron transfer integral and hence the bandwidth is small [60, 63-65]. In the<br />

high temperature limit this gives a factor of T 3/2 instead of a T in the drift and<br />

Hall mobilities. There are some unphysical assumptions required for the<br />

non-adiabatic analysis to be valid. The assumptions behind the adiabatic


Electronic and Magnetic Measurements 35<br />

polaron formation are more physically plausible and therefore the adiabatic<br />

analysis is preferred.<br />

3.1.3.2.3 Diffusive Conductivity<br />

The same form of the conductivity derived for adiabatic small polaron<br />

hopping is found more generally for diffusion limited conduction. If the<br />

charge carrier must overcome an activation energy, E A , to hop to a<br />

neighboring site, the probability for hopping will be proportional to<br />

exp(-E A /k B T). From the theory of the random walk, the diffusion constant D<br />

can be estimated using this hopping probability, the frequency, ω, with which<br />

an attempt to hop is made (usually the frequency of the optical phonon which<br />

provides the lowest barrier to hopping at some instant), and the site to site<br />

distance, a: D = λωa 2 exp(-E A /k B T). λ is a geometrical factor approximately<br />

equal to 1. The mobility is related to the diffusivity via the Nernst-Einstein<br />

relation μ = eD/k B T. Thus the conductivity, σ = neμ, of a general, activated,<br />

diffusive process is σ = ne 2 ωa 2 /k B T -1 exp(-E A /k B T). The transport of ions in a<br />

crystal is diffusive, and therefore the ionic conductivity is often analyzed<br />

assuming this form of the conductivity [66]. The transport of small polarons<br />

is very similar and has also been described in this way [67-69].<br />

3.1.3.2.4 Variable range Hopping<br />

For a semiconductor or insulator at low enough temperatures, the<br />

predominant conduction mechanism may no longer be by excitation of<br />

carriers to the mobility edge or by thermally activated hopping to the nearest<br />

neighbor but by variable range hopping. At low temperatures, the<br />

mechanism with the lowest barrier energy will dominate. Due to any<br />

randomness in the sample, the hopping site with the lowest barrier energy<br />

will not in general be the nearest neighbor. The increased hopping distance<br />

will of course reduce the probability that the carrier will tunnel to this


36 Chapter 3<br />

position; however, this is always offset by the lower barrier energy at<br />

sufficiently low temperatures.<br />

The simplest quantitative derivation of the form of variable range<br />

hopping is the following. For a given site, the number of states within a<br />

range R per unit energy is (4π/3)R 3 N(E F ), where N(E F ) is the density of<br />

localized states. Thus the smallest energy difference for a site within a radius<br />

R is on average the reciprocal of this ΔE = 1/[(4π/3)R 3 N(E F )]. Thus, the further<br />

the carrier hops, the lower the activation energy.<br />

The carrier has an electronic wave function exponentially localized on a<br />

particular site with a decay or localization length of ξ. The tunneling<br />

probability that the electron will hop to a site a distance R away will contain a<br />

factor exp(-2R/ξ). The further the distance, the lower the tunneling<br />

probability.<br />

Since the hopping favors large R while the tunneling favors small R,<br />

there will be an optimum hopping distance R for which the hopping<br />

probability proportional to exp(-2R/ξ) exp(ΔE/k B T) is a maximum. This will<br />

occur when 1/R 4 = 8πN(E F )k B T/ξ. Substituting this value for R, the hopping<br />

probability and thus the conductivity is proportional to exp(-(T 0 /T) 1/4 ) where<br />

T 0 = Cξ 3 /k B N(E F ). C is a constant which in this derivation is 24/π ≈ 7.6, but<br />

other values of C are obtained from more sophisticated analyses. C ≈ 21 is<br />

recommended by Shklovskii and Efros.<br />

The conductivity for variable range hopping is usually given the form σ 0<br />

exp(-(T 0 /T) ν ), ν = 1/4. The exact form of σ 0 depends on the model and may<br />

have a power law temperature dependence of its own. For example T 0.33 has<br />

been found [70].<br />

Other values of ν can be obtained theoretically using different<br />

assumptions. The above derivation assumes a three dimensional system. In


Electronic and Magnetic Measurements 37<br />

general, ν = 1/d+1 where d is the dimension of the system. In particular this<br />

gives ν = 1/2 and ν = 1/3 for one and two dimensional systems respectively.<br />

The exponent ν = 1/2 is also found when the Coulomb interaction between<br />

the electrons is taken into account [71]. This interaction produces a gap in the<br />

density of states. If this gap is larger than the bandwidth of the hopping<br />

carriers, ν = 1/2 is found. If polarons are the charge carriers, the range of<br />

validity for ν = 1/d+1 may be reduced to lower temperatures than normally<br />

expected [72].<br />

3. 1. 3. 3 Poor Metals / Heavily doped semiconductors<br />

Some materials which would normally be considered metals, have<br />

dρ/dT < 0 over a significant temperature range. Amorphous metals and<br />

metallic glasses, which can be prepared by rapid cooling, film deposition, or<br />

irradiation, for example, can show a resistivity which decreases slightly with<br />

temperature. Most examples are transition metal alloys with zero<br />

temperature resistivities approaching the Ioffe-Regel limit, where the mean<br />

free path is equal to the interatomic distance, corresponding to about 1 mΩcm<br />

[69]. Heavily doped semiconductors can show similar behavior described<br />

below.<br />

The physics behind this phenomenon is not completely understood, and<br />

the data is usually interpreted on a case by case basis. The theoretical<br />

conductivity is generally described as a power series with respect to<br />

temperature where the power ranges from zero to one, e.g. σ = σ 0 + mT n ,<br />

n < 1, and m can be positive or negative. n = 1/2 or 1 is often observed and<br />

can be explained theoretically. For example, Altshuler and Aronov predict an<br />

n = 1/2 correction for the effect of long-range interaction between electrons<br />

[70]. For semiconductors doped near the metal-insulator transition, n = 1/3<br />

has been observed and explained theoretically [70].


38 Chapter 3<br />

Complex materials often have resistivities that are inadequately described<br />

by the simple models given above. For example the cuprate superconductors<br />

and SrRuO 3 show a linear resistivity well above the Ioffe Regel limit [73]. A<br />

more extreme example is Ba 6 Co 25 S 27 which not only has a resistivity<br />

minimum but a less than linear resistivity above the minimum [42]. As<br />

more materials with complex electronic structures are examined, such non-<br />

standard behavior will certainly become more common.<br />

3. 1. 3. 4 Phase transitions<br />

Any precise measurement as a function of temperature can usually detect<br />

or be influenced by a phase transition. Conductivity data in particular can be<br />

quite precise but can also be greatly influenced by subtle electronic, magnetic<br />

or structural changes. Thus one should be cautious when fitting data to very<br />

similar functional forms. Electronic transitions such as metal-insulator<br />

transitions and charge ordering (charge density wave formation) obviously<br />

change the conductivity at the ordering temperature. Magnetic transitions<br />

are also easily observed in the conductivity since local moments act as<br />

scattering centers, or may even help localize/delocalize carriers.<br />

Ferromagnetic metals for instance always show a decrease in resistivity as the<br />

temperature goes below the Curie temperature [73]. Structural phase<br />

transitions usually change the symmetry and volume, which affects the<br />

conduction paths and density of the charge carriers. This will subtlety change<br />

the conductivity at the transition temperature even if the conduction<br />

mechanism remains the same, as shown in section 4.2.5.<br />

Careful measurements can help determine the nature of such phase<br />

transitions. If the phase transition is reversible and not hysteretic, i.e. the data<br />

are the same upon warming and cooling, then the phase transition is<br />

probably a single second order process. If the data is hysteretic then a first<br />

order process is involved. For example, the ferromagnetic and accompanying<br />

metal insulator transition in the manganites appears to be of second order.


Electronic and Magnetic Measurements 39<br />

The charge ordering transition in the nonmagnetic manganites is also not<br />

hysteretic and probably second order. However, in samples where both<br />

ferromagnetism and charge ordering occur, the conductivity and magnetic<br />

susceptibility are hysteretic, indicating a more complex transition. In this<br />

case, there is no subgroup-supergroup relationship between the two ordered<br />

phases; so to go from the magnetically ordered phase to the charge ordered<br />

phase requires a first order (hysteretic) phase transition [74, 75].<br />

In the region near a second order phase transition (critical regime) physical<br />

quantities such as electronic transport are often best described by a power law<br />

about the critical point, e.g. (T - T C ) n where T C is the critical temperature and n<br />

is the critical exponent. For example, the resistivity of a ferromagnet near the<br />

Curie temperature should be described by a critical exponent (α) equal to that<br />

of the heat capacity [73]. Magnetic critical exponents are described in more<br />

detail in section 3.2.2.2.4<br />

3.2 Magnetism<br />

The magnetism of a material is usually dominated by the electrons near<br />

the Fermi level. Unpaired, localized electrons in insulators have very<br />

different magnetism than paired or itinerant ones. Thus, magnetic<br />

measurements can detect the subtle electronic structure of a material. Since<br />

these measurement do not require contacts, they can be quite easy to perform<br />

Ð and have made them quite standard in characterizing new materials.<br />

3. 2. 1 Measurement<br />

There are various ways to measure a magnetic field and thus<br />

magnetization. The most common instruments to measure D.C.<br />

magnetization are the vibrating sample magnetometer, Faraday balance and<br />

SQUID magnetometer. Surface magnetization can be accurately measured by<br />

the Magneto-optical Kerr effect (MOKE). The apparatus used in this study is a<br />

Quantum Design MPMSR 2 SQUID magnetometer.


40 Chapter 3<br />

3.2.1.1 Apparatus<br />

The Quantum Design SQUID (Superconducting Quantum Interference<br />

Device) MPMS (Magnetic Property Measurement System) is a commercially<br />

available Òturn-keyÓ magnetometer. The device runs on liquid He and can<br />

reach temperatures of 2 K to 400 K and fields up to 7 Tesla. A sample heater is<br />

available to achieve temperatures 300 K to 800 K but the sample space<br />

diameter is reduced from 8 mm to about 3 mm. The revision 2 (MPMSR 2 )<br />

improves the software, allowing simple programming for measurements<br />

lasting several days without requiring maintenance.<br />

The SQUID coils detect the longitudinal component of the magnetization<br />

as the sample is pulled through them. The coils are wound in a second-<br />

derivative configuration in which the upper and lower single turns are<br />

counter wound with respect to the two-turn center coil. This configuration<br />

strongly rejects interference from nearby magnetic sources and allows the<br />

system to function without benefit of a superconducting shield around the<br />

SQUID sensing loops. The raw data are a set of voltage readings taken as a<br />

function of position as the sample is moved upward throughout the sensing<br />

loops. The data are normally fit to a theoretical signal to calculate the<br />

magnetic moment. Several vertical scans are averaged to obtain a standard<br />

deviation.<br />

The two important data reduction algorithms are compared in Figure 3-1<br />

for a standard Yttrium Iron Garnet (YIG) sample. The Òlinear regressionÓ<br />

assumes the sample is properly centered. If the sample is not exactly centered<br />

(longitudinally), the measured moment will be noticeable different. The<br />

Òiterative regressionÓ iterativly finds the center, so if the sample is off-center<br />

the reported moment does not change. When doing temperature scans, the<br />

length of the sample rod will significantly change. The iterative regression is<br />

clearly superior for these measurements. The precision of these two<br />

algorithms are the same (Figure 3-1). The Òfull scanÓ algorithm is simply the


Electronic and Magnetic Measurements 41<br />

sum of the squares of the raw data and does not fit the data to any particular<br />

curve.<br />

Moment (emu)<br />

0.852<br />

0.850<br />

0.848<br />

0.846<br />

0.844<br />

0.842<br />

Scan Length and Algorithm<br />

Iterative<br />

Linear<br />

Full Scan<br />

0.840<br />

2 3 4 5 6 7 8<br />

Scan Length (cm)<br />

Figure 3-1 Effect of changing the scan length for<br />

magnetization measurements. A YIG crystal is used. The<br />

three data reduction schemes are also compared. The results<br />

for the full scan algorithm with scan length less than 5 cm are<br />

off scale.<br />

The distance the sample is pulled through the coils, called the scan length,<br />

can be varied. The effect of the scan length is shown in Figure 3-1. Long scan<br />

lengths are not ideal since field and temperature gradients exist in the sample<br />

chamber. Very long (> 6 cm) will even destabilize the chamber temperature,<br />

requiring several minutes to stabilize before each scan. Field gradients will<br />

cause irreversible flux motion in superconductors - changing the magnetic<br />

moment. Short scans can drastically reduce the accuracy of the moment<br />

without affecting the precision (Figure 3-1). Ferromagnetic and paramagnetic<br />

substances are not very irreversibly affected by field gradients, so a medium<br />

scan length of about 5 cm is used to ensure good accuracy.


42 Chapter 3<br />

The second-derivative configuration of the SQUID coils will, in principle,<br />

not pick up the magnetization of the sample holder if it has a uniform<br />

magnetization. For this reason, a plastic straw with the sample wedged in the<br />

center makes an ideal sample holder. Shrink wrap tubing, shrunk around<br />

the sample also makes a good, but less rigid, sample holder. If string or grease<br />

is used to hold the sample, then there will be an additional contribution to<br />

the magnetization due to the sample holder. Similarly, if the sample is<br />

between two quartz rods, there will be a paramagnetic contribution to the<br />

magnetization due to the gap between the diamagnetic quartz. Materials not<br />

containing transition elements have a diamagnetic susceptibility of about<br />

-0.5 × 10 -7 emu/g, so light sample holders (plastic) are somewhat preferable to<br />

heavier ones (quartz glass).<br />

The measured moment is also somewhat affected by the horizontal<br />

centering of the sample. Rotating the sample along the field axis should have<br />

no effect on the moment. However, it may very slightly change the<br />

horizontal position of the sample. For whatever reason, the moment can<br />

change a few percent with the sample rotator. Such a sensitivity on<br />

orientation probably results in some of the unexpected jumps in the data.<br />

The YIG sample has a moment 96.6% of that expected. Before an<br />

experiment the sample is rotated to give the maximum moment. This is to<br />

minimize the possible underestimation of the moment and usually provides<br />

a region where the moment is less sensitive to small changes of the rotation.<br />

The transverse SQUID coils can in principle also measure the net<br />

transverse magnetic moment. A longitudinal component to the<br />

magnetization will affect the transverse measurement. The sample can be<br />

rotated along the axis of the sample rod, which in combination with the<br />

transverse measurements can be used to measure the transverse-only<br />

moment.


Moment (emu)<br />

0.234<br />

0.233<br />

0.232<br />

0.231<br />

0.230<br />

0.229<br />

0.228<br />

Electronic and Magnetic Measurements 43<br />

Sample Rotation<br />

0.227<br />

0 50 100 150 200 250 300 350<br />

Rotation Angle (°)<br />

Figure 3-2 Magnetization of YIG sample as it is rotated along<br />

the field axis.<br />

3.2.2 Analysis<br />

The important quantities in magnetic measurements are the magnetic<br />

field vector H, measured in Oersted and the magnetization intensity (or<br />

simply the magnetization) M, measured in emu/volume (electromagnetic<br />

unit = erg/Oersted). M is often reported in units of emu/g, emu/mole or<br />

Bohr magnetons (μ B ) per magnetic atom. The magnetic induction field B<br />

measured in gauss, defined by B = H + 4π M, is often considered more<br />

fundamental in physics. The above units are CGS units which are more<br />

prevalent than MKS since the equations tend to be simpler. The MKS unit<br />

which is used commonly in the magnetic literature is the Tesla = 10000 gauss.<br />

Most materials have a magnetization proportional to the applied field.<br />

The magnetic susceptibility χ, defined by M = χH, relates these two quantities.<br />

It is convenient to use χ = M/H, for non-ferromagnetic and isotropic


44 Chapter 3<br />

substances. Otherwise, M = M 0 + χH + χ 3 H 3 + É can be used where χ is now a<br />

second rank tensor describing the linear response, M 0 is the spontaneous<br />

magnetization, and χ n describe the higher order nonlinear terms.<br />

The magnetic field H can be produced either by electric currents or by the<br />

magnetization. If a sample is magnetized, an H field is produced in the<br />

sample to oppose the magnetization and therefore demagnetize the sample.<br />

This demagnetization field H d is for an ellipsoidal sample directly<br />

proportional to the magnetization H d = N d M, where 0 ≤ N d ≤ 4π is the<br />

demagnetization factor. Since the demagnetization field depends on the<br />

shape of the sample, it is the internal field H i = H a - H d the applied field minus<br />

the demagnetization field which affects the magnetic response. Since the<br />

correction for the demagnetization fields is often ignored, one should be wary<br />

of the effects of demagnetization. Ferromagnets in small applied fields for<br />

example may have reduced bulk magnetization due to demagnetization.<br />

3. 2. 2. 1 Diamagnetism and Paramagnetism<br />

Since atoms are made up of charged particles which undergo orbital<br />

motion and have magnetic moments, there are many ways a materials can<br />

have a magnetic response. A diamagnetic response is that for which the<br />

susceptibility is negative while a positive contribution to the susceptibility is<br />

paramagnetic.<br />

3.2.2.1.1 Larmor diamagnetism<br />

Since the electrons in an atom are essentially free charges orbiting a<br />

nucleus, the application of a magnetic field, by LenzÕs law, induces an<br />

opposing magnetic moment. The resulting magnetic susceptibility is<br />

therefore negative and known as Larmor or core-electron diamagnetism.<br />

Both the classical and quantum mechanical analyses yield the same result,<br />

namely that the diamagnetic susceptibility of an atom or ion is proportional<br />

to the number of electrons it contains, Z and its cross sectional area


Susceptibility (10 -6 emu/G g)<br />

0.0<br />

-0.1<br />

-0.2<br />

-0.3<br />

Electronic and Magnetic Measurements 45<br />

χ mol = -ZN A e 2 /6mc 2 where is the mean square electron radius. Since<br />

atomic sizes are roughly the same, and masses nearly proportional to Z, the<br />

diamagnetic susceptibility per gram for all substances are about the same<br />

χ g ≈ -0.5 × 10 -6 emu/g G. Since it is difficult to calculate accurately, it is<br />

best to use experimental values (tabulated in [76, 77]). In most measurements<br />

of paramagnetic and ferromagnetic substances, the diamagnetism is so small<br />

it is usually ignored.<br />

An example of where core diamagnetism is important is shown in<br />

Figure 3-3. Oxide substrates are generally diamagnetic, but provide a<br />

significant susceptibility due to their large mass compared to the mass of a<br />

thin film. The diamagnetism is fairly temperature independent but<br />

substrates may contain paramagnetic impurities which alter the low<br />

temperature diamagnetism.<br />

Diamagnetic Susceptibility of Substrates<br />

LaAlO 3<br />

Al 2 O 3<br />

MgO<br />

-0.4<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

Figure 3-3 Diamagnetic magnetic susceptibility of typical<br />

substrates. The increase in the susceptibility at low<br />

temperatures is due to paramagnetic impurities.


46 Chapter 3<br />

3.2.2.1.2 Conduction electron diamagnetism<br />

The motion of conduction electrons in a metal or semiconductor will also<br />

provide a diamagnetic response (Landau diamagnetism) to a magnetic field.<br />

This is difficult to calculate but generally of the same order as the Pauli<br />

paramagnetism (section 3.2.2.1.3). A superconductor, however, expels a<br />

magnetic field completely (Meissner effect) by the motion of the<br />

superconducting electrons. The diamagnetism of a superconductor is,<br />

therefore, very large, χ vol = -1/4π emu/cc G in the Meissner regime. Since<br />

superconductivity is affected by temperature and a magnetic field, the<br />

magnetic response of a superconductor is actually quite complicated.<br />

3.2.2.1.3 Pauli paramagnetism<br />

Electrons in a metal can be partitioned into spin-up and spin-down bands,<br />

parallel and antiparallel to an applied magnetic field H. The magnetic field<br />

will lower the energy of the spin-up band compared to the spin-down band<br />

(by 2μ B H) and spin-down electrons will flip their spins and pour over into the<br />

spin-up band. The number of electrons (per volume) that need to flip their<br />

spins is approximately the density of electronic states, n(E F ) times one half of<br />

the energy splitting. This produces a net magnetization proportional to the<br />

2<br />

magnetic field and therefore a positive susceptibility, χvol = μB n(EF ). A more<br />

sophisticated statistical-mechanical derivation produces the same result with<br />

small correction proportional to T 2 . The Pauli susceptibility of a metal should<br />

thus be nearly temperature independent and about the same magnitude as<br />

the Larmor diamagnetism. Since the contribution of the Landau<br />

diamagnetism is not known, it is usually not possible to get more than an<br />

order-of-magnitude estimate of the Pauli susceptibility from magnetization<br />

measurements. In principle, however, the Pauli susceptibility should be a<br />

measure of the bare density of electronic states at the Fermi level, n(E F ), in the<br />

absence of many body effects. The linear electronic specific heat term


Electronic and Magnetic Measurements 47<br />

(γ, section 3.3.2.1) is also proportinal to n(E F ), as well as the electronic effective<br />

mass. The ratio of these two experimental results is known as the Wilson<br />

ratio.<br />

3.2.2.1.4 Curie paramagnetism<br />

If a material has unpaired electrons, their magnetic spins will be highly<br />

susceptible to a magnetic field. The energy to orient the spins comes from the<br />

magnetic energy MH which for an atom with effective magnetic moment p, is<br />

μ B pH. The thermal energy k B T, randomizes the spins to oppose the<br />

alignment, so the amount of alignment is roughly proportional to μ B pH/k B T.<br />

The net magnetization is then μ B p times the fraction of moments that are<br />

aligned. So, the magnetic susceptibility of free spins should behave like<br />

2 2<br />

μB p /kBT. This is several hundred times larger at room temperature than the<br />

core or Pauli diamagnetism/paramagnetism, and significantly temperature<br />

dependent.<br />

The full statistical mechanical treatment for quantum mechanical<br />

moments with spin S, orbital angular momentum L, and total angular<br />

momentum J, gives a magnetization per atom M = gμ B JB J (gμ B JH/k B T). g is the<br />

LandŽ g-factor, which for atoms where J = S is the just the electron g-factor,<br />

3 1 ⎡SS<br />

( + 1) − LL ( + 1)<br />

⎤<br />

g = 2.00, otherwise gJLS ( , , ) = +<br />

2 2<br />

⎢<br />

⎣ JJ ( + )<br />

⎥.<br />

1 ⎦<br />

Transition metal ions generally have their orbital angular momentum<br />

quenched so J ≈ S. The Brillouin function, B J (x), is defined by<br />

2J+ 1 x( 2J + 1)<br />

1 x<br />

J + 1<br />

BJ ( x)<br />

= coth − coth . For small x, BJ ( x)≈x , while for<br />

2J<br />

2J<br />

2J 2 J<br />

3J<br />

very large x, BJ (x) ≈ 1. The magnetization is then linear with respect to the<br />

magnetic field until it approaches its saturation value of gμ B J. In the linear<br />

region, μ B H « k B T, the susceptibility has the Curie form χ mol<br />

2 2<br />

NAμ<br />

B gJJ ( + 1)<br />

=<br />

3k<br />

T<br />

B


48 Chapter 3<br />

Susceptibility (emu/G mol)<br />

4.0 10 -2<br />

3.5 10 -2<br />

3.0 10 -2<br />

2.5 10 -2<br />

2.0 10 -2<br />

1.5 10 -2<br />

1.0 10 -2<br />

5.0 10 -3<br />

0.0 10 0<br />

where N<br />

2<br />

AμB 1 emu K<br />

≈<br />

. Since the spin state is not necessarily known, the<br />

3kB<br />

8 Gauss mol<br />

effective moment p is often reported where p 2 = g 2 J(J + 1). The Curie<br />

susceptibility diverges as the temperature approaches zero; however, in real<br />

systems the spins will Òfreeze-outÓ below some temperature making, for<br />

example, an antiferromagnet, ferromagnet, or spin glass.<br />

Magnetization (μ B /Fe)<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

-1.0<br />

-1.5<br />

-80 -60 -40 -20 0 20 40 60 80<br />

Magnetic Field (kOe)<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

Figure 3-4 Paramagnetic susceptibility and hysteresis loop of a<br />

paramagnetic Fe containing organometallic compound [78].<br />

Insulators with magnetic atoms far apart (do not interact) are good<br />

examples of Curie paramagnets. Figure 3-4 shows the magnetic properties of<br />

an organometallic compound [SC(CH3 ) 2C(CH3 )NCH2CH2NH2 ] -<br />

2 FeCl with well<br />

isolated Fe +3 ions [78]. The magnetization is slightly nonlinear at 5 K, as<br />

predicted by the Brillouin function. The linear susceptibility can be fit to<br />

χ = χ 0 + C/T where χ 0 is the diamagnetic correction (section 3.2.2.1.1). If the


Electronic and Magnetic Measurements 49<br />

diamagnetic correction is assumed to be known, and the effective Curie<br />

moment as a function of temperature can be extracted. The compound<br />

[SC(CH3 ) 2C(CH3 )NCH2CH2CH2 ] 2N - Fe PF -<br />

6 in Figure 3-5 has a gradual spintransition<br />

of the Fe +3 from S = 1/2 to S = 5/2 as the temperature is increased<br />

[78].<br />

Effective paramagnetic moment (μ B )<br />

5.5<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

Figure 3-5 Effective paramagnetic moment of Fe in the<br />

organometallic compound [SC(CH3 ) 2C(CH3 )NCH2CH2CH2 ] 2N -<br />

showing a spin transition [78].<br />

FePF -<br />

6<br />

3.2.2.2 Ferromagnetism<br />

If the magnetic species interact, then their magnetic properties can be<br />

profoundly altered. Neighboring moments which prefer to be aligned<br />

parallel lead to ferromagnetism while antiparallel alignment gives<br />

antiferromagnetism. More complex interactions lead to ferrimagnetism,<br />

canted ferromagnetism, and spin glass behavior.


50 Chapter 3<br />

H/M (T/μ B )<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

5T<br />

5000G<br />

5Gauss<br />

Calculated 1/χ for SrRuO 3<br />

0<br />

160 180 200 220 240 260 280 300<br />

Temperature (K)<br />

Figure 3-6 Calculated inverse magnetic susceptibility of<br />

SrRuO 3 using the molecular field model.<br />

3.2.2.2.1 Weiss molecular field model<br />

The phenomena of ferromagnetism, antiferromagnetism and<br />

ferrimagnetism can often be well described by a mean field or molecular field<br />

model. The molecular field model simply assumes that all the interactions<br />

from the neighboring magnetic species can be described in terms of an<br />

effective internal or molecular field H m , which is proportional to the magnetic<br />

order H m = λM (for convenience in this section, M will refer to the<br />

magnetization per molecule). For a ferromagnet λ > 0. The total magnetic<br />

field on an atom is then the sum of the applied field H a and the molecular<br />

field. Substituting H a + H m = H + λM for H gives a transcendental function of<br />

M in terms of H and T which can be solved iteratively. For a simple<br />

ferromagnet the only parameter is the molecular field constant λ, which can<br />

2 2<br />

be determined from the critical temperature TC via λ = 3kBTC /μB p .


Electronic and Magnetic Measurements 51<br />

Above the Curie temperature the molecular field model gives the Curie-<br />

Weiss law for the zero-field susceptibility χ mol<br />

2 2<br />

NAμ<br />

B gJJ ( + 1)<br />

=<br />

3k<br />

T − Θ , where ΘP is<br />

B P<br />

the paramagnetic Curie temperature. In the Curie-Weiss molecular field<br />

model Θ P = T C . Experimentally, for reasons discussed below, the experimental<br />

value of Θ P is usually a slightly larger than the true critical temperature T C .<br />

In a field, this susceptibility is slightly less than linear (χ 3 < 0), like it is for<br />

paramagnetic moments, due to saturation. The calculated inverse<br />

susceptibility, 1/χ = H/M, is shown in Figure 3-6 for SrRuO 3 . The 1/χ = 0<br />

intercept of the extrapolation of 1/χ vs. T is the paramagnetic Curie<br />

temperature Θ P .<br />

Below T C the mean field magnetization decreases exponentially slowly as<br />

the temperature is raised from T = 0 K, as shown in Figure 3-7. The zero field<br />

magnetization decreases to zero at the ferromagnetic Curie temperature T C .<br />

Also at low temperatures, the mean field magnetic susceptibility is very<br />

small, i.e. the application of a magnetic field increases the calculated<br />

magnetization only slightly (Figure 3-7).


52 Chapter 3<br />

Magnetization (μ B )<br />

Calculated Mean Field Magnetization for SrRuO<br />

3<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Magnetization (μ B )<br />

2.005<br />

2.000<br />

1.995<br />

1.990<br />

0.0<br />

0 50 100<br />

Temperature (K)<br />

150 200<br />

The molecular field model assumes very little about the microscopic<br />

origin of ferromagnetism, only that it makes sense to define a local molecular<br />

field. Microscopic models for the interaction of atomic magnetic moments<br />

can be formulated into a similar mean field model. In this case, the<br />

interatomic exchange energy J can be related to the molecular field constant λ.<br />

A microscopic model also leads to a more accurate prediction of the<br />

magnetization. At low temperatures, the quantized moments can be<br />

collectively excited as discussed in section 3.2.2.2.8 to give the faster decrease<br />

in the magnetization than the molecular field model. The microscopic<br />

model also gives a more realistic description of the observed properties near<br />

the critical point (section 3.2.2.2.4).<br />

Mean Field Forced Magnetization<br />

5K<br />

50K<br />

1.985<br />

0 20 40 60 80 100<br />

Magnetic Field (kOe)<br />

5T<br />

5kOe<br />

5 Oe<br />

Figure 3-7 Mean field magnetization calculated in various<br />

fields for SrRuO 3 with T C = 165K. Inset show the very small<br />

field dependence of the magnetization (forced magnetization)<br />

in this model.


Electronic and Magnetic Measurements 53<br />

3.2.2.2.2 Itinerant electron Model<br />

The ground state properties of a metallic ferromagnet are believed to be<br />

fairly well described by the band theory which has advanced in the recent<br />

decades [79]. The thermal excitations and finite temperature properties,<br />

however, are more difficult because of the effects of strong electron-electron<br />

correlations in narrow energy bands. This limits the utility of traditional<br />

Hartree-Fock or mean field theory described briefly in this section and<br />

requires a treatment of generalized spin density fluctuations described i n<br />

section 3.2.2.2.3 to describe fully the magnetic excitations in itinerant electron<br />

ferromagnets.<br />

The quantum mechanical exchange interaction can lead to<br />

ferromagnetism in an electron gas within the band model, commonly known<br />

as Stoner ferromagnetism [80, 81] or itinerant electron ferromagnetism.<br />

Using the molecular field approximation, a ferromagnetic state is realized for<br />

Jn(E F ) > 1 where J is the average intra-atomic exchange energy per atom and<br />

n(E F ) is the density of states at the Fermi level. This is also known as the<br />

Stoner condition for ferromagnetism. For Jn(E F ) < 1 the material is not<br />

ferromagnetic but has a Pauli susceptibility enhanced by 1/(1-Jn(E F )) [82]. This<br />

factor is called the Stoner enhancement factor.<br />

The effect of the magnetization at finite temperatures can be calculated<br />

using the mean field approximation. Specifically, the energy of the up (down)<br />

spin band is uniformly lowered (raised) by an amount proportional to the<br />

magnetic polarization ζ and a molecular field constant. Since this model<br />

assumes electrons are either in up or down spin bands, the excitations are<br />

known as single particle, spin flip, or Stoner excitations. There is no energy<br />

gap required for Stoner excitations. However, zero or low energy excitations<br />

require a non-zero momentum transfer, while excitations with zero<br />

momentum transfer require an excitation energy equal to the exchange<br />

splitting as can be seen in Figure 3-8. The region of allowed Stoner excitations


54 Chapter 3<br />

is called the Stoner continuum. Collective, spin-wave excitations are not (yet)<br />

included. The mean field analysis [81] gives (M/M(0,0)) 2 = 1-(T/T C ) 2 +<br />

2χ(0,0)H/M.<br />

As expected from a Ginzberg-Landau/mean-field approximation section<br />

3.2.2.2.5, this formula predicts mean field exponents of β = 1/2 and γ = γ ′ = 1<br />

for T near T C , and linear M 2 vs. H/M Arrott plots. Above T C the inverse<br />

susceptibility is proportional to T 2 2<br />

-TC [81], which is in gross disagreement<br />

with the experimentally observed Curie-Weiss law (1/χ = T - T C ). The finite<br />

zero field susceptibility at T = 0, χ(0,0), is essentially the enhanced Pauli-<br />

susceptibility [83].<br />

Unlike the local moment model which requires the zero temperature<br />

saturation moment M(0,0) to be equal to the sum of all the moments, the<br />

band model allows for opposing moments in the minority band at T = 0 (ζ


Electronic and Magnetic Measurements 55<br />

T/T C « ζ 2 (fractional polarization)[81]. A more complete improvement of the<br />

Stoner theory takes into account the effect of spin density fluctuations. Such<br />

improvements are discussed the following section.<br />

3.2.2.2.3 Generalized Model<br />

The self-consistent-renormalization (SCR) theory [79, 84, 85] combines<br />

both single particle and spin wave excitations into generalized interacting<br />

spin density fluctuations. It deals with the coupled modes of spatially<br />

extended spin density fluctuations in a self-consistent fashion. The theory<br />

gives a new mechanism for the Curie-Weiss magnetic susceptibility<br />

associated with the spatially extended modes of spin fluctuations instead of<br />

the local moments as spatially localized spin fluctuations. Furthermore, the<br />

theory is successful quantitatively in explaining various physical properties of<br />

weak itinerant ferro- and antiferromagnets: the theory contains a limited<br />

number of parameters which can be determined experimentally. Thus by the<br />

1970Õs the theory of magnetism was advanced to the stage of having two<br />

established regimes in the opposite extremes, the local moment systems and<br />

weak itinerant ferro- and antiferromagnets. In the former regime the spin<br />

fluctuations are localized in real space while in the latter they are localized in<br />

k-space. In order to unify the two extremes into a single theory an<br />

interpolation theory for the intermediate regime was developed in the 1980Õs<br />

using the functional integral method within the static approximation.


56 Chapter 3<br />

Energy<br />

Magnetic Excitation Spectrum<br />

Spin<br />

Waves<br />

Stoner<br />

Continuum<br />

Momentum Transfer<br />

Figure 3-8 Energy spectrum of magnetic excitations. Spin<br />

wave excitations have a one-to-one dispersion relation while<br />

excitations in the Stoner continuum (shaded region) do not.<br />

The intensity of excitations in the Stoner continuum is<br />

strongest where the spin waves meet the continuum.<br />

The theory predicts that spin waves exist only in a small region at the<br />

origin of (ω,k) space with dispersion relation ω = k 2 D, where D is the spin<br />

wave stiffness. The strongest intensity of the dynamical susceptibility is in<br />

the region where the spin wave excitations meet the Stoner continuum<br />

shown in Figure 3-8. In the low temperature limit a T 3/2 law is predicted i n<br />

the magnetization. However, calculations for Ni 3 Al [86] show that the<br />

magnetization is best approximated by a T 2 dependence over a broad<br />

temperature range below T C . It has been assumed that in the highly<br />

correlated limit, i.e. metallic ferromagnets with a large saturation moment,<br />

the magnetization will obey the spin wave T 3/2 law [85], presumably because<br />

of strong evidence for the existence of spin waves in metallic ferromagnets.


Electronic and Magnetic Measurements 57<br />

The major advancement of the SCR theory is the development of a new<br />

mechanism for the Curie-Weiss magnetic susceptibility applicable to itinerant<br />

electron ferromagnets. The paramagnetic Curie-Weiss susceptibility arises<br />

essentially from the opposition of the magnetic energy to order the spins and<br />

the thermal energy to disorder them. The magnetization is then proportional<br />

to µ B H/k B T, giving the inverse susceptibility H/M linear in T. In the case of<br />

the weak itinerant electron ferromagnets there are no localized spins above<br />

T C so the susceptibility cannot arise from this mechanism. Instead, the<br />

changing mean-square local amplitude of the spin fluctuation provides the<br />

Curie-Weiss susceptibility.<br />

The SCR expression for the inverse magnetic susceptibility contains a<br />

2<br />

contribution from the mean-square local amplitude of the spin fluctuation SL [84] which can be derived from the fluctuation-dissipation theorem.<br />

2<br />

Calculations have shown that SL increases almost linearly with temperature<br />

above T C giving rise to the Curie-Weiss law. This is in contrast with the local<br />

2<br />

moment mechanism where SL is a constant. The Curie constant of the new<br />

mechanism is related to the band structure around the Fermi surface and is<br />

independent of the saturation moment at T = 0. This Curie-Weiss law should<br />

hold even for paramagnetic metals when they are very close to the<br />

ferromagnetic instability, in contrast to that predicted by paramagnon theories<br />

[84].<br />

2<br />

The temperature variation of SL (T), which determines the new Curie-<br />

Weiss mechanism is strongest when the stiffness of the longitudinal spin<br />

fluctuations is small. When the longitudinal stiffness is small, a relatively<br />

2<br />

rapid increase in SL (T) with inverse temperature is expected above TC and it<br />

should saturate at a certain value determined by the band structure. After<br />

saturation the spin fluctuations behave like local moments with a certain<br />

degree of short range order, since the local amplitude of the spin density is<br />

almost fixed. This phenomenon is called temperature-induced local


58 Chapter 3<br />

moments [84]. Thus, a subtle change is expected in the Curie-Weiss<br />

mechanism at some temperature above T C . This has been used to explain the<br />

change in the Curie constant at around 3T C in Co(S/Se) 2 [87]. This<br />

phenomenon is also expected in some exchange-enhanced paramagnets.<br />

In Heusler Alloys such as Pd 2 MnSn, Ni 2 MnSn, Cu 2 MnAl, etc., manganese<br />

atoms are spatially well separated from each other (> 4Å), and are believed to<br />

carry well-defined local moments since the Rhodes-Wohlfarth ratio is close to<br />

one and neutron scattering experiments can be well analyzed with the<br />

Heisenberg model. The ferromagnetic MnPt 3 , FePd 3 and antiferromagnetic<br />

FePt 3 are likely to have well-defined 3d local moments that couple with<br />

itinerant 4d or 5d electrons of Pd or Pt. CoS 2 and the metallic ferromagnets<br />

AMnO 3 described in chapter 4 fit perhaps best in this category.<br />

In the intermediate regime between the local moment ferromagnets and<br />

the weak itinerant-electron ferromagnets are the Ferromagnetic transition<br />

metals (Fe, Co, and Ni) and compounds Fe 3 Pt and CeFe 2 . These materials<br />

have relatively high T C ’s but have Rhodes-Wohlfarth ratios [84, 88] greater<br />

than one. The ferromagnetic oxide SrRuO 3 also fits in this category as<br />

described in detail in Appendix A.<br />

MnSi is a good example of a weak-ferromagnetic helimagnet which is well<br />

described by the SCR theory. Other weak ferromagnets such as Ni 3 Al, ZrZn 2<br />

and Sc 3 In can also be described in terms of the SCR theory [84].<br />

Metallic Cr which orders antiferromagnetically below 312K with a long<br />

period spin density wave, is the most often used example of an itinerant<br />

electron antiferromagnet. However, the Curie-Weiss law is not obeyed above<br />

T N . The magnetism of Cr is best described with an itinerant electron approach<br />

with the addition of a nesting Fermi surface model to describe the spin<br />

density waves [84]. The antiferromagnetic metals γ-Mn, γ-(FeMn) and γ-Fe<br />

have also been studied [84].


Electronic and Magnetic Measurements 59<br />

Examples of nearly ferro- or antiferromagnetic metals which show a large<br />

susceptibility are CoSe 2 (Curie Weiss) Pd or YCo 2 [84], and V 2 O 3 [89]. The<br />

nearly ferro- or antiferromagnetic semiconductor FeSi (Curie Weiss) shows a<br />

Curie-Weiss susceptibility above 700K with a broad transition at 500K to a low<br />

susceptibility state [84].<br />

3.2.2.2.4 Critical region<br />

As the temperature approaches the critical temperature T C , the<br />

spontaneous magnetization vanishes and the susceptibility diverges. This<br />

makes these properteis not analytic near T C . The observed properties are well<br />

described by a power law |T - T C | n where n is a scaling exponent. For<br />

example, the spontaneous magnetization varies as (T C -T) β for T < T C , while<br />

the susceptibility varies as (T C - T) γ for T > T C , and for T = T C (critical isotherm)<br />

M δ = H. The scaling exponents in the critical region are called critical<br />

exponents.<br />

The critical exponents are very similar for a wide range of second order<br />

phase transitions. They tend to depend only on the general form of the<br />

interaction causing the phase transition, such as the dimensionality, and not<br />

on the particular material. Thus most ferromagnets fall into the universality<br />

class of 3-dimentional Heisenberg or Ising ferromagnets. The theoretical<br />

critical exponents calculated for these models is given in Table 3-1. There are<br />

only two unique critical exponents in each universality class. The others are<br />

related via the scaling relations such as γ = β(δ -1).<br />

According to the scaling hypothesis [90], the magnetic equation of state in<br />

the critical region depends only on the scaled variables H/|T C /T - 1| β+γ and<br />

M/|T C /T - 1| β . For example, a plot of the scaled M 2 vs. the scaled H/M, (scaled


60 Chapter 3<br />

Arrott plot), will then have only two curves: one branch for the T < T C data<br />

and another for T > T C .<br />

The magnetization near T C is predicted in the SCR theory, described i n<br />

4/3 4/3 1/2<br />

section 3.2.2.2.3, to behave as M = (TC -T ) [84, 91] which reducecs to β = 1<br />

for T near T C .<br />

Table 3-1 Theoretical 3-dimentional critical exponents for<br />

different models and selected experimental values [92, 93].<br />

β γ δ<br />

Ising .33 1.24 4.8<br />

Heisenberg .36 1.39 4.8<br />

Mean Field .5 1 3<br />

ZrZn 2 .50(3) 1.02(5) 3.1(3)<br />

Fe, Ni, YIG .37(2) 1.2(2) 4(1)<br />

3.2.2.2.5 Landau mean field theory<br />

Near the critical temperature T C the molecular field, or mean field model<br />

(section 3.2.2.2.5) predicts mean field critical exponents (Table 3-1). The<br />

Landau theory of continuous, second order phase transitions (excluding<br />

fluctuations) arrives at the same mean field result. Here the free energy is<br />

expanded in a Taylor series of the order parameter (M in the case of<br />

ferromagnetism). Due to the symmetry of the order parameter, only even<br />

powers of M are nonzero. The first few terms are [94]: G = G 0 + a(T - T C )M 2 +<br />

bM 4 - HM. At a given H and T, M can be found by minimizing the free energy<br />

G. The general solution is H/M = 2a(T - T C ) + 4bM 2 . Below T C the saturation<br />

magnetization (H = 0) is found to be M 2 = (T C - T)a/2b, giving the critical<br />

exponent β = 1/2. Above T C in a field, M is small so the bM 4 term can be<br />

ignored. This gives a susceptibility χ = M/H = (T - T C ) -1 /2a, and a critical


Electronic and Magnetic Measurements 61<br />

exponent γ = 1. Along the critical isotherm T = T C , H = 4bM 3 so the critical<br />

exponent δ = 3.<br />

3.2.2.2.6 Arrott Plot<br />

According to the Landau equation of state H/M = 2a(T - T C ) + 4bM 2 , a plot<br />

of M 2 vs. H/M at a constant temperature gives a straight line. At T C , this line<br />

intersects the origin. The M 2 = 0 intercept gives the inverse susceptibility<br />

1/χ 0 (H = 0) while the H/M intercept gives the spontaneous magnetization<br />

2<br />

M0 (H = 0).<br />

In real systems near the critical point, non mean field critical exponents<br />

are observed. Thus the plot of M 2 vs. H/M (Arrott plot) will not necessarily<br />

be straight lines, but allows a visual analysis of the data [94]. The T C<br />

determined from the Arrott plot is the isotherm that extrapolates to the<br />

origin. Likewise, the extrapolation to the axis intercepts give the H = 0<br />

spontaneous magnetization M 0 (T < T C ) and inverse susceptibility<br />

1/χ 0 (T > T C ) and allows a visual exstimation of the error. From these M 0 (T)<br />

and χ 0 (T) the critical exponents β and γ can be estimated by fitting<br />

M 0 (T) ∝ (1 - T/T C ) β and χ 0 (T) ∝ (T/T C - 1) -γ .<br />

If the assumption of scaling is incorrect, such as a crossover from one<br />

scaling region to another as the temperature changes, then the entire data will<br />

not scale with the same exponents. In this case, the method described above<br />

is better for characterizing the separate regions. Such a crossover is observed<br />

in SrRuO 3 where β appears to change from 0.32 to 0.38 (Appendix A) so that<br />

the entire data does not scale very well with a single average value for β.<br />

3.2.2.2.7 The Curie temperature<br />

Because the Curie temperature affects many of the properties, there are<br />

several ways of measuring it. One of the simplest is finding the paramagnetic


62 Chapter 3<br />

Curie temperature, the temperature where the linear extrapolation of the 1/χ<br />

curve crosses 1/χ = 0. Near the critical temperature however, 1/χ is not<br />

necessarily linear in T since 1/χ = (T - T C ) γ where γ is usually large than 1.<br />

Thus the paramagnetic Curie temperature is usually slightly larger than the<br />

true critical temperature.<br />

In principle the thermodynamic critical temperature can be found by<br />

fitting the divergence of all the measured quantities and using the scaling<br />

relations. This assumes a true critical region has been reached where the<br />

critical exponents do not change. The simpler method using the Arrot plots,<br />

described in the previous section (3.2.2.2.6) can give as accurate results. Thus,<br />

the Arrott T C is probably the best simple way of finding the critical<br />

temperature.<br />

The most basic definition of T C for a ferromagnet is the temperature where<br />

the spontaneous magnetization drops to zero (as the temperature is raised).<br />

Unfortunately due to demagnetization effects, the zero applied field<br />

magnetization will drop to zero at T C faster and less uniformly than expected.<br />

Nevertheless, the temperature where the magnetization, measured in zero or<br />

small applied field, is often used to measure the critical temperature and is<br />

called here the remnant T C .<br />

In an applied field, the magnetization does not drop to zero because of the<br />

large susceptibility near T C . However, below T C the magnetization has a<br />

negative curvature while above T C it has a positive curvature. Thus the<br />

inflection point of the M vs. T curve (denoted here as the inflection T C ) is<br />

often used to measure the Curie temperature. In the mean field model, the<br />

inflection T C is quite close to the real critical temperature. Figure 3-7 shows<br />

the calculated M vs. T of SrRuO 3 in various fields. In a 5 T applied field, the<br />

inflection point is 1.6 K higher than the true T C used for the calculation. This<br />

agrees well with the experimental data shown in Figure 3-9, where the


Electronic and Magnetic Measurements 63<br />

derivatives are calculated numerically from the measured magnetization. In<br />

some systems, the inflection T C can increase substantially as the applied field<br />

is increased, as seen for Gd 0.67 Ca 0.33 MnO 3 in section 5.1.4, making it a poor<br />

estimate of the Curie temperature.<br />

Magnetization (µ B )<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

SrRuO 3 Inflection T C<br />

163 K<br />

167 K<br />

166 K<br />

164 K<br />

5.5T<br />

3T<br />

1T<br />

5kOe<br />

1kOe<br />

300 Oe<br />

50 Oe<br />

0.0<br />

140 150 160 170 180 190<br />

Temperature (K)<br />

Figure 3-9 The inflection T C measured for a SrRuO 3 pellet.<br />

For H < 1 Tesla the inflection T C is within 1 K of the Arrott T C<br />

= 163 K. At higher H the inflection T C increases by only a few<br />

degrees.<br />

3.2.2.2.8 Spin waves<br />

Collective excitations (spin waves) are expected to be the dominant mode<br />

of excitations at the lowest temperatures. This is because it can be shown that<br />

collective excitations will always have lower energy than single spin<br />

excitations. The spin wave theory has been mostly developed for localized<br />

moments [95] but can also be generalized to include k-space electrons [81].


64 Chapter 3<br />

The energy ω of a spin wave is given by ω = Dk 2 + ∆ for low momentum k<br />

spin waves. D is the spin wave stiffness and ∆ is a gap energy arising from<br />

anisotropy or applied magnetic fields. With no gap the spontaneous<br />

magnetization decreases as T 3/2 [95]: M(T)/M 0 = 1 - (T/Θ 3/2 ) 3/2 . For a<br />

3 4<br />

=<br />

2. 612<br />

⎛ nS ⎞ πD<br />

. The<br />

⎝ ⎠ kB specific heat of these excitations [95] also follows a T 3/2 law:<br />

ferromagnet with spin S and spin density n, Θ3/ 2<br />

c V = 1.925 k B nS(T/Θ 3/2 ) 3/2 . The spin wave stiffness D may be estimated from<br />

the critical temperature T C and number of near neighbors z, using D = 2JS 5/3<br />

and k B T C /J = 0.0521(z - 1)(11S(S + 1)-1) [96], where J is the exchange energy.<br />

Correction factor M(t)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

0.0 0.5 1.0 1.5 2.0<br />

0.0<br />

0 2 4 6 8 10<br />

k B T/gµ B H<br />

Figure 3-10 Correction factor to the T 3/2 contribution of the<br />

magnetization in the spin wave theory due to a magnetic field<br />

H.<br />

The first order effect of a magnetic field providing a gap ∆ = gµ B H /k B T, is<br />

to adjust the T 3/2 coefficient by a function of k B T/gµ B H: M(T)/M 0 =<br />

1 - F 3/2 (k B T/gµ B H)/ζ(3/2) (T/Θ 3/2 ) 3/2 . This correction factor M(t), shown in<br />

2


Electronic and Magnetic Measurements 65<br />

Figure 3-10, depends on the temperature T, thus altering the temperature<br />

dependence of the magnetization to have an effective power higher than 3/2<br />

[97]. The Bose-Einstein integral function [95] F P (x) is given by F P (x) =<br />

and ζ(P) =<br />

−<br />

e<br />

n<br />

∞ n x<br />

∑ P<br />

n=<br />

/<br />

1<br />

∞<br />

1<br />

∑ (e.g. ζ(3/2) = 2.612 and ζ(1/2) = -1.460). The correction C(t) to<br />

1 n P<br />

n=<br />

the T 3/2 contribution to the heat capacity is similar: C(t) = [F 5/2 (t) + 4F 3/2 (t)/5t +<br />

4F 1/2 (t)/15t 2 )]/ζ(5/2), where t = k B T/gµ B H. The correction factor C(t) is shown<br />

in Figure 3-11. Due to the Maxwell relation the increase in the effective<br />

power in M(T) also increases the effective power of dc/dH (T).<br />

Correction Factor C(t)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.0 0.5 1.0 1.5 2.0<br />

0.0<br />

0 2 4 6 8 10<br />

k B T/gµ B H<br />

Figure 3-11 Correction factor to the T 3/2 contribution of the<br />

heat capacity in the spin wave theory due to a magnetic field<br />

H.


66 Chapter 3<br />

3.2.2.2.9 Irreversibility<br />

Below T C , a ferromagnet must undergo a first order phase transition when<br />

the field direction is changed. Like all first order processes, this involves<br />

nucleation and growth (of magnetic domains) and can be hysteretic.<br />

Technically there is only hysteresis when the internal field changes sign; if<br />

the field increases or decreases while retaining the same sign (direction), the<br />

change in the magnetization is continuous. However, in real systems, there<br />

exist internal anisotropy and demagnetization fields. So in low fields<br />

irreversible behavior can be found.<br />

Magnetization (µ B )<br />

0.12<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

SrRuO 3 Pellet irreversibility<br />

Field Cooled<br />

20 Oersted Magnetization<br />

Zero Field Cooled<br />

FC<br />

ZFC<br />

0.00<br />

0 50 100<br />

Temperature (K)<br />

150 200<br />

Figure 3-12 SrRuO 3 showing spin-glass like irreversibility of<br />

zero-field-cooled and field-cooled measurements in a small<br />

field. The field cooled curve may look saturated, but is<br />

actually less than 1/10 saturated at low temperatures. A small<br />

peak is observed in the zero-field-cooled measurement when<br />

the reversibility point is reached.<br />

The field required to reduce to zero the magnetization of a saturated<br />

sample, is called the coercive field. When the coercive field is large, such as


Electronic and Magnetic Measurements 67<br />

in polycrystalline SrRuO 3 , the irreversibility will be evident in many<br />

measurements. The coercivity depends on the microstructure of the sample,<br />

and therefore depends on the processing. SrRuO 3 single crystal for instance<br />

(Appendix A) has a coercive field 100 times smaller than polycrystalline<br />

SrRuO 3 .<br />

Magnetization (µ B )<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

SrRuO 3 initial magnetization<br />

Inflection Point<br />

0 1000 2000 3000 4000 5000<br />

Magnetic Field (Gauss)<br />

Figure 3-13 Initial magnetization of SrRuO 3 pellet at 5 K, after<br />

cooling in zero field. The magnetization follows a “S” shaped<br />

curve providing an inflection point.<br />

In the irreversible regime, a ferromagnet will have magnetic properties<br />

similar to that of a spin glass [98]. For example, it is common to characterize<br />

spin glasses [99] (or disordered superconductors) by comparing the field-cooled<br />

(sample cooled below T C in the measuring field) and zero-field-cooled<br />

(sample cooled in zero field and then the field applied before measuring upon<br />

warming) magnetization in a small field. For a spin-glass or disordered<br />

superconductor, these two curves are different. In such small fields, a


68 Chapter 3<br />

ferromagnet will also show such hysteresis, as demonstrated in Figure 3-12<br />

for SrRuO 3 . The field cooled curve may look like a ferromagnetic M vs. T<br />

curve, however it is influenced greatly by the demagnetization field and<br />

coercivity. For example in applied fields lower than the maximum<br />

demagnetization field (4πNM S ) the net magnetization can only increase until<br />

the demagnetization field cancels the applied field (M = H/4πN). The a plot<br />

of M vs. T in this case will show a constant, low value for M, until the true<br />

magnetization decreases near T C (see for example the low field data in [29]).<br />

Reversible Field (H rev ) (Gauss)<br />

100000<br />

10000<br />

1000<br />

100<br />

SrRuO 3 Irreversibility Line<br />

Hrev (const T)<br />

Hrev (Const H)<br />

H rev = (1-T/T C ) 1.5<br />

10<br />

0.01 0.1<br />

1-T/Tc<br />

1<br />

Figure 3-14 Magnetic irreversibility line for polycrystalline<br />

SrRuO 3 . Above the line the magnetization is reversible,<br />

below it is irreversible. The irreversibility exponent is about<br />

1.5.<br />

Another characteristic of a spin glass is an “S” shaped magnetization<br />

curve, which gives an inflection point in M vs. H. Such a curve is also<br />

observed for a ferromagnet, such as that seen for SrRuO 3 in Figure 3-13. The<br />

curves for increasing and decreasing fields will meet at the irreversibility<br />

point. The irreversibility points measured by the two methods (constant T,


Electronic and Magnetic Measurements 69<br />

Figure 3-13; and constant H, Figure 3-12) give an irreversibility line, much<br />

like that in superconductors. For polycrystalline SrRuO 3 the reversible field is<br />

approximately proportional to (1 - T/T C ) 1.5 .<br />

Like many glassy systems, the magnetic properties in the irreversible<br />

region are time dependent. In a ferromagnet the domain walls are pinned<br />

and so must overcome some barrier energy to move. Thus the<br />

magnetization increases in a time dependent manner. Figure 3-15 shows the<br />

magnetization of a SrRuO 3 pellet at 5 K after increasing the magnetic field<br />

from 0 Gauss (where the magnetization was 0.0076 µ B ) to 100 Gauss. The<br />

magnetization increases with a Log(time) dependence, with 98% of the<br />

increase happening before the first data point (about 1 minute). Such a<br />

logarithmic time dependence is also characteristic of a spin glass [99].<br />

SrRuO 3 has all the properties of a ferromagnet, and because of this long<br />

range order can not be a spin glass. The domain structure of a ferromagnet<br />

provides magnetic disorder on a larger length scale than a spin glass, but to be<br />

a spin glass, the disorder must be on the atomic scale.<br />

3.2.2.3 Antiferromagnetism<br />

In an antiferromagnetic substance, the neighboring magnetic moments<br />

align antiparallel. Such a material is called an antiferromagnet. In the<br />

ordered state, there is often no net moment in zero field. Thus magnetization<br />

measurements cannot easily measure the magnetic order. In neutron<br />

diffraction, the magnetic unit cell multiplies giving extra diffraction peaks.<br />

The intensity of these peaks is a direct measure of the ordering.<br />

In a magnetic field, an antiferromagnet has a small positive susceptibility.<br />

Above the ordering temperature, or Néel temperature T N , the susceptibility<br />

can be well described by the Curie-Weiss law used for ferromagnets. Since the<br />

interactions are antiparallel, the coupling is of the opposite sign as that for a


70 Chapter 3<br />

Magnetization (µ B )<br />

0.01645<br />

0.01640<br />

0.01635<br />

0.01630<br />

0.01625<br />

0.01620<br />

Magnetization Creep<br />

100 1000<br />

Time (s)<br />

10000<br />

Figure 3-15 Time dependent magnetization of SrRuO 3 pellet<br />

at 5 K. The field was increased from 0 to 100 Gauss. The<br />

magnetization follows a Log(time) dependence.<br />

ferromagnet. This leads to a negative paramagnetic Curie temperature<br />

Θ P ≈ -T N , where the susceptibility above T N is given by χ mol<br />

2 2<br />

NAµ<br />

B gJJ ( + 1)<br />

=<br />

3k<br />

T − Θ .<br />

B P<br />

Below the ordering temperature T N , M vs. H is still linear with no H = 0<br />

magnetization, but the susceptibility depends on the orientation of the<br />

magnetic field with respect to that of the magnetic moments [56]. For<br />

example, Figure 3-16 shows the susceptibility of platinum containing<br />

“CaRuO 3 ” crystals, or CaRu/PtO 3 described in section 2.1.2.1. If the magnetic<br />

field is applied parallel to the ordered moments, the applied torque (µ × H) is<br />

small so the susceptibility decreases to zero. The CaRu/PtO 3 crystal when<br />

oriented with the 2-fold symmetric axis parallel to H shows such a<br />

susceptibility. When the field is perpendicular to the ordered moments, the<br />

torque is maximized providing a large, constant susceptibility in the ordered


χ (emu/g/G)<br />

4.0 10 -5<br />

3.5 10 -5<br />

3.0 10 -5<br />

2.5 10 -5<br />

2.0 10 -5<br />

1.5 10 -5<br />

1.0 10 -5<br />

Electronic and Magnetic Measurements 71<br />

5.0 10 -6<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

state. This is seen in the crystals oriented with a 3-fold symmetric axis parallel<br />

to H. Without Pt, CaRuO 3 has an antiferromagnetic like susceptibility<br />

(1/(T + Θ P )) above 50K (Θ P ≈ -210K p eff ≈ 9.75 or S ≈ 1.14) but does not become<br />

antiferromagnetic. Instead, it remains paramagnetic but with a stronger<br />

temperature dependence ≈ 1/T (Figure 3-16). From 2 K to 20 K the<br />

susceptibility fits well to (1/(T + Θ P )) with an antiferromagnetic slope<br />

(Θ P = -73K p eff = 4.17 or S = 0.64).<br />

CaRuO 3 and Pt Containing CaRuO 3 Crystal<br />

CaRuO 3<br />

3-fold axis crystal<br />

2-fold axis crystal<br />

Figure 3-16 Magnetic susceptibility of a Pt containing<br />

“CaRuO 3 ” crystal. The crystal was aligned with its 2-fold<br />

symmetric axis parallel to the applied field has a susceptibility<br />

characteristic of antiferromagnetic moments aligning parallel<br />

to the field, while the 3-fold axis appears to have moments<br />

perpendicular to the field.<br />

3.2.2.4 Ferrimagnetism<br />

If the sublattices of different moment orientation do not exactly cancel,<br />

there is then a small ferromagnetic moment in the ordered state, that behaves


72 Chapter 3<br />

with temperature and field much like a ferromagnet. When this is due to a<br />

canting of neighboring moments from truly parallel or antiparallel, this is<br />

called a canted antiferromagnet. If the ferromagnetic moment is not due<br />

entirely to canting but arises from the different strengths of the sublattice<br />

magnetizations, the material is called a ferrimagnet.<br />

Most commercial insulating “ferromagnets” are actually ferrimagnets.<br />

Common examples are ferrites and garnets such as YIG. The multiple<br />

magnetic interactions inherent to a ferrimagnet provide a variety a magnetic<br />

behavior. For example, when the magnetizations of the opposing sublattices<br />

exactly cancel, the net moment is zero like it is in an antiferromagnet. The<br />

temperature where this occurs is called the compensation temperature T Comp .<br />

3.2.2.4.1 Mean field model for Gd 0.67Ca 0.33MnO 3<br />

The gross features of a ferrimagnet can be explained using the molecular<br />

field model of section 3.2.2.2.5. For a ferrimagnet with two sublattices, the<br />

total magnetization M is the sum of the two sublattice magnetizations M A and<br />

M B . The molecular field acting on each sublattice can then arise from<br />

interactions with atoms in the other sublattice as well as atoms in the same<br />

sublattice. For example on the A sublattice H mA = λ AA M A + λ AB M B . The<br />

molecular field constants λ ij can be either positive for ferromagnetic coupling<br />

between sublattices or negative for antiferromagnetic coupling. For a<br />

ferrimagnet or antiferromagnet, there must be strong antiferromagnetic<br />

coupling. Similarly, on the B sublattice H mB = λ BB M B + λ AB M A . Once the<br />

molecular field constants are known, the magnetization of each sublattice can<br />

be calculated iteratively as described in section 3.2.2.2.5 until self consistency is<br />

reached.


Gd 0.7Ca 0.3MnO 3<br />

Electronic and Magnetic Measurements 73<br />

Hwang et al.<br />

Figure 3-17 Temperature - tolerance factor phase diagram<br />

from reference [100], with the position of Gd 0.67 Ca 0.33 MnO 3<br />

indicated.<br />

This has been done for Gd 0.67 Ca 0.33 MnO 3 where the experimental results<br />

are given in chapter 5. In Gd 0.67 Ca 0.33 MnO 3 the Mn atoms are on one<br />

sublattice (B) and Gd atoms on the other (A). The dominant molecular field<br />

constant is expected to be the ferromagnetic Mn-Mn coupling due to the<br />

stronger 3d exchange interaction. From the trend found in the manganese<br />

ferromagnetic ordering temperature as a function of the average of the R and<br />

A atom size [100] in R 1-x A x MnO 3 , shown in Figure 3-17, the T C is expected to be<br />

around 60K. The molecular field constant can then be approximated using<br />

the relationship found for a ferromagnet in section 3.2.2.2.5, namely<br />

2 2<br />

λBB = 3kBTC /µ B<br />

pB .<br />

2<br />

Here pi also includes the site occupancy for each type of atom on the i<br />

sublattice. For Gd0.67Ca0.33MnO3 , the Mn (B) sublattice has two magnetic<br />

contributions since the site is 67% occupied by Mn +3 (S = 2) and 33% occupied


74 Chapter 3<br />

by Mn +4 2<br />

(S = 3/2). This gives pB = Σ nig 2 si (si +1) = 21. On the Gd/Ca (A)<br />

sublattice, only Gd +3 2<br />

(67% occupied, S = 7/2) has a moment giving pA = 42.<br />

The Gd-Gd interaction is expected to be negligible, and therefore λ AA ≈ 0,<br />

since the Gd 4f electrons are well localized. The Gd atoms are expected to<br />

order due to the significant Gd-Mn interaction which is antiferromagnetic, so<br />

λ BA < 0. In the case of Gd 0.67 Ca 0.33 MnO 3 , there exists a temperature where the<br />

magnetization of the larger, but poorly ordered, Gd moments exactly cancels<br />

that of the Mn sublattice. This temperature is called the compensation<br />

temperature T Comp .<br />

The compensation temperature T Comp can be used to estimate the<br />

molecular field constant λ BA . The weakly coupled Gd approximately act like<br />

free paramagnetic spins reacting to the internal (molecular) field caused by<br />

the ordered Mn moments. In zero applied field this gives<br />

2 2<br />

MA = HmA µ B<br />

pA /3kBT. The internal field on the A sublattice is related to λBA from H mA = λ BA M B . At T = T Comp , there is no net magnetization so M B + M A = 0.<br />

2 2<br />

Combining these relations gives λBA = -3kBTComp /µ B<br />

pA .<br />

Above TN , the susceptibility can be calculated analytically [101] giving a<br />

hyperbola for 1/χ vs. T shown in Figure 5-5. The high temperature asymptote<br />

is a straight, Curie-Weiss like, line which intersects the T axis at<br />

2 4 2 2 2 2<br />

ΘP = (µ B /3kB ) (pB λBB + 2pB pA λBA )/(pB + pA ). For Gd0.67Ca0.33MnO3 , λBA < 0 and<br />

2<br />

λBB > |λBA | so ΘP should be less than but close to 7λBB (µ B /3kB ) which is 1/3 of<br />

the value that would be expected (TC of Mn) if no Gd moments were present.<br />

In this model, the ferrimagnetic Néel temperature is given by<br />

2 2<br />

TN = pB λBB (µ B /3kB )[1/2 + √(1/4 + (pAλBA /pBλBB ) 2 )] [101]. For λBB > |λBA | this is<br />

2 2<br />

only slightly greater than the TC = pB λBB (µ B /3kB ) expected if the Gd moments<br />

were absent. Combining this result with that of the last paragraph gives a


Electronic and Magnetic Measurements 75<br />

2 2 2<br />

general requirement of this model TN /ΘP > (pB + pA )/pB = 3, which should be<br />

easily verified from experiments (section 5.1.5.1).<br />

Magnetization 2 (µ B /mole) 2<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Calculated Gd 0.67 Ca 0.33 MnO 3 Arrott Plot<br />

80<br />

83K<br />

85K<br />

0<br />

0 2000 4000 6000 8000 10000 12000<br />

H/M (Oe/µ B )<br />

Figure 3-18 Calculated Arrott Plot for Gd 0.67 Ca 0.33 MnO 3 using<br />

the mean field model with T C = 83.3 K.<br />

The critical temperature T C , of a ferrimagnet is the Néel temperature T N .<br />

As T approaches T N from below, the zero field magnetization vanishes much<br />

like that of a ferromagnet. The theoretical Arrott plot T C (shown in Figure 3-<br />

18) is equivalent to the remnant T C = T N . The mean field calculation gives a<br />

nearly linear critical isotherm on the Arrott plot, shown in Figure 3-18, since<br />

M δ = H and δ = 3 is the mean field critical exponent. The inflection point in<br />

the M vs.. T curve is also a good estimate of T C . In this model, the inflection<br />

T C increases by only 0.55 K/Tesla as H is increased.<br />

Below T C the calculated magnetic susceptibility is small since the internal<br />

molecular field is quite large. The ferrimagnet susceptibility calculated here,<br />

however, is not uniformly decreasing as T decreases like it does for a


76 Chapter 3<br />

ferromagnet. Instead, it peaks as shown in Figure 3-19 near the compensation<br />

temperature T Comp .<br />

Susceptibility (µ B /T)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Calculated Susceptibility of Gd 0.67 Ca 0.33 MnO 3<br />

χ = (M(6T) - M(4T))/2T<br />

0<br />

0 20 40 60 80 100<br />

Temperature (K)<br />

Figure 3-19 High field differential susceptibility for<br />

Gd 0.67 Ca 0.33 MnO 3 calculated using the mean field model. The<br />

maximum is at 11.5 K which is near T Comp = 14.2 K in this<br />

model.<br />

3.3 Heat Capacity<br />

The heat capacity of a sample, particularly at low temperatures, is a<br />

powerful experimental technique which gives fundamental information<br />

about low-lying excitations of many kinds (phonons, electrons, magnons,<br />

etc.). Microscopic models which describe these excitations typically give<br />

quantitative predictions of the specific heat which can then be used to check<br />

the theory. The has been quite successful at low temperatures where the<br />

excitations are simple. At higher temperatures, the dispersion relation of the<br />

excitation spectrum can become less simple and therefore complicate the heat


Electronic and Magnetic Measurements 77<br />

capacity. Thermal measurements at higher temperatures T>100K are typically<br />

only used to study phase transitions.<br />

3. 3. 1 Measurement<br />

The most straightforward method of measuring heat capacity is adiabatic<br />

calorimetry. Here, the heat capacity C=∆Q/∆T is measured directly by applying<br />

a known ∆Q and measuring ∆T. This technique requires that the thermal<br />

relaxation time constant between the sample and its surroundings be large<br />

compared to the measurement time. In a common Differential Scanning<br />

Calorimeter (DSC) or Differential Thermal Analyzer (DTA) the loss of heat<br />

due to thermal relaxation is accounted for by subtracting the signal of an<br />

empty sample holder measured at the same rate. A DSC actually operates by<br />

recording the heat input required ∆Q /dt to keep the temperature changing at<br />

a constant rate ∆T/dt = constant. The DTA supplies a constant ∆Q and<br />

measures ∆T between the sample and a reference. These commercial<br />

instruments are used mostly to detect and estimate the entropy associated<br />

with a phase transition. Very accurate measurements require a more<br />

complicated apparatus with better thermal isolation.<br />

The relaxation technique and ac methods for measuring heat capacity are<br />

preferred when measuring a small ∆Q or ∆T. The heat ∆Q is actually<br />

calculated in the relaxation technique by measuring the thermal time<br />

constant of the system. The sample is attached to an object with high thermal<br />

mass and known temperature (the surroundings) by means of a weak<br />

thermal link. After heating the sample, the temperature of the sample<br />

exponentially decays to that of the surroundings. The heat capacity is then<br />

the product of the time constant of this decay and the thermal conductance of<br />

the weak link. This thermal conductivity can be measured in the steady state<br />

by applying a fixed power and measuring the resulting temperature of the<br />

sample.


78 Chapter 3<br />

3.3.1.1 Apparatus<br />

The apparatus, method of operation, data reduction and analysis used at<br />

Stanford is described in detail in [102] and references therein. The sample is<br />

attached to a bolometer by means of Wakefield thermal compound. The<br />

bolometer is a patterned sapphire substrate containing two phosphorus doped<br />

silicon thermometers (used for different temperature ranges) and a heater<br />

with aluminum contact pads (6). Gold-7%Cu wire of 0.001” diameter is<br />

spotwelded to these contacts to mechanically, electrically, and thermally<br />

connect the sample containing bolometer to the surrounding sample holder.<br />

The gold wire provides the weak thermal link to the surroundings and has a<br />

thermal conductivity approximately linear in temperature. The sample<br />

chamber is evacuated before the measurement and care must be taken to keep<br />

out He since it adds to the thermal conductivity and even the heat capacity if<br />

it condenses on the bolometer. The heat capacity due to the sapphire<br />

bolometer, thermal compound, gold wire and aluminum contacts (in order of<br />

decreasing importance) must be subtracted from the raw heat capacity data.<br />

The precision of the measurement is typically 0.5% with or without an<br />

applied magnetic field.<br />

3.3.2 Analysis<br />

The specific heat at low temperatures is usually of the form c = γT + βT 3 .<br />

Because of this, the data are usually plotted as c/T vs. T 2 giving a straight line<br />

with slope β and intercept γ.<br />

The specific heat of spin waves are discussed in section 3.2.2.2.8.<br />

3. 3. 2. 1 Electronic specific heat<br />

The linear term comes from the specific heat of itinerant electrons and is<br />

proportional to the density of electronic states at the Fermi level, n(E F ).<br />

2 2<br />

γ = n(EF )kB π /3. Insulators and semiconductors have no free electrons and<br />

therefore no linear term. Good metals tend to have broad bands with low


Electronic and Magnetic Measurements 79<br />

density of states, giving a γ of the order 1 mJ/mol/K 2 . Bad metals tend to<br />

have narrow bands and high density of states providing the region of large γ<br />

between semiconductors and metals.<br />

3. 3. 2. 2 Phonon specific heat<br />

The cubic term in the specific heat arises from the excitation of phonons in<br />

the crystal lattice, and can be related to the Debye temperature Θ D :<br />

-3<br />

β = ΘD nkB 12π 4 /5. The Debye temperature varies like the spring constant<br />

between neighboring atoms in the crystal: with Θ D being very large for light,<br />

strongly bonded atoms such as diamond, and small for heavy, weakly bonded<br />

atoms such as lead metal. Graphite, which has strong bonding in only two<br />

dimensions, has a specific heat over a significant temperature range which is<br />

proportional to T 2 instead of T 3 .<br />

The Debye model does not account for the specific heat of optical phonons.<br />

At higher temperatures this may become important. Using the Einstein<br />

model the additional specific heat is proportional to nk<br />

Θ E is the Einstein temperature.<br />

B<br />

2 Θ E ( ΘE<br />

/ T) e<br />

Θ E / T<br />

( e − 1)<br />

/ T<br />

2<br />

where


4. Intrinsic Electrical Transport and Magnetic Properties of<br />

La0.67Ca0.33MnO3 and La0.67Sr0.33MnO3 MOCVD Thin Films and<br />

Bulk Material<br />

In this study, measurements among polycrystalline pellets, single crystals<br />

and thin films, supplemented with literature data when available, are<br />

compared to establish the properties inherent to La 0.67 AE 0.33 MnO 3 where AE =<br />

Ca, Sr, Ba or Pb. Whereas most reported La 0.67 AE 0.33 MnO 3 films have been<br />

grown by laser ablation, it is shown here that organometallic chemical vapor<br />

deposition (MOCVD) can produce high quality films. It will be demonstrated<br />

that singular behavior can be obtained in bulk and thin films. Most of the<br />

work described in this section has been previously published [103].<br />

To form the thermodynamically stable phase, the material is annealed at<br />

high temperatures for long times and cooled slowly. This produces films,<br />

polycrystalline pellets and single crystals with very similar properties.<br />

Properties which are common to all these samples are called “intrinsic” to the<br />

thermodynamically stable phase. The remaining differences can then be<br />

attributed to inhomogeneities, noncrystallinity, microstructure, strain or<br />

growth induced effects. Major changes in the magnetization and electrical<br />

transport have been observed in RE 0.67 AE 0.33 MnO 3+δ compounds (where RE is<br />

a rare earth element) when the oxygen stoichiometry was varied; these have<br />

been discussed mainly in terms of non-stoichiometry doping and carrier<br />

localization [14-17, 104, 105].<br />

In order to compare experimental properties certain terms need to be<br />

defined. For simplicity, T C the used in this chapter is the temperature where<br />

the bulk of the zero field magnetization disappears. T MI (metal to insulator<br />

temperature) is defined here as the temperature where the resistivity is a<br />

maximum. T MR is the temperature where the magnetoresistance (R(0)-R(H))<br />

80


Magnetization (emu/g)<br />

80<br />

60<br />

40<br />

20<br />

Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

Magnetization (emu/g)<br />

100<br />

50<br />

0<br />

-50<br />

La 0.67 Sr 0.33 MnO 3 Polycrystalline Pellet<br />

0<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

is the largest. The term “CMR” is used here when the magnetoresistance<br />

ratio ∆R/R(H) is greater than 10 (or 1000%).<br />

b<br />

5K<br />

-100<br />

-75 -50 -25 0 25 50 75<br />

Magnetic Field (kOersted)<br />

10 kOersted<br />

4.1 Magnetism<br />

The polycrystalline samples have relatively square hysteresis loops (Figure<br />

4-1), with forced magnetization at 70kOe of only a few percent. The saturation<br />

magnetizations are close to that expected for high spin manganese in<br />

octahedral coordination: for spin only (orbital contribution quenched)<br />

moment µ = gsµ b , g=2 and then µ= 2µ b [0.67 x 2 (from Mn 3+ ) + 0.33 x 3/2 (from<br />

Mn 4+ )] = 3.67µ b . The measured ferromagnetic and Curie temperatures are the<br />

same within experimental error of a few degrees. The physical properties of<br />

the polycrystalline materials summarized in Table 4-1 Physical Properties of<br />

Polycrystalline Pellets are consistent with previous results [1-4, 29, 106, 107].<br />

10<br />

8<br />

6<br />

4<br />

2<br />

a<br />

100 Oe<br />

0<br />

370 372 374 376378 380<br />

Temperature (K)<br />

Figure 4-1 Magnetization of La 0.67 Sr 0.33 MnO 3 polycrystalline<br />

pellet at 10kOe. Inset a, magnetization at 100Oe used to<br />

determine T C = 375K. Inset b, full hysteresis loop at 5 K.<br />

81


82 Chapter 4<br />

M/M 0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Since the precise oxygen concentration has a noticeable effect on T C and M S<br />

[104], slight differences are expected. Float zone grown crystals have similar<br />

magnetic properties.<br />

M/M 0<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

La 0.67 Sr 0.33 MnO 3 Film<br />

-1.0<br />

-75 -50 -25 0 25 50 75<br />

5 kOersted<br />

0.0<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

Table 4-1 Physical Properties of Polycrystalline Pellets<br />

TC MS a<br />

La0.67Sr0.33MnO3 376K<br />

3.59µb /Mn<br />

3.88Å<br />

La0.67Ca0.33MnO3 270K<br />

3.39µ b /Mn<br />

3.86Å<br />

HC 20Oersted 10Oersted<br />

Curie Weiss µ eff<br />

Hall carrier density (of films)<br />

5.61µ b<br />

2.1holes/Mn<br />

5.96µ b<br />

0.9holes/Mn<br />

5K<br />

Magnetic Field (kOersted)<br />

Figure 4-2 Magnetization of La 0.67 Sr 0.33 MnO 3 film (LSM1) on<br />

LaAlO 3 at 5kOe. Inset, full hysteresis loop at 5 K of film and<br />

(diamagnetic) substrate.


Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

The transition temperatures of the annealed films are close to those of the<br />

polycrystalline material, and are summarized in Table 4-2. The saturation<br />

magnetization, measured at 5kOe, also decreases as T 2 for both samples. The<br />

annealed films exhibit a (5 K) coercivity of about 100 Oe and a sheared<br />

hysteresis loop expected from demagnetization or uniaxial anisotropy effects<br />

(Figure 4-2 and Figure 4-3).<br />

M/M 0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

M/M 0<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

La 0.67 Ca 0.33 MnO 3 Film<br />

5K<br />

-1.0<br />

-75 -50 -25 0 25 50 75<br />

Magnetic Field (kOersted)<br />

5 kOersted<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

Figure 4-3 Magnetization of La 0.67 Ca 0.33 MnO 3 film (LCM15) on<br />

LaAlO 3 at 5kOe. Inset, full hysteresis loop at 5 K of film and<br />

(diamagnetic) substrate.<br />

4. 1. 1 Low Temperature Excitations<br />

The saturation magnetization decreases approximately proportional to T 2<br />

at low temperatures (Figure 4-4). This temperature dependence of M can be<br />

expected for metallic ferromagnets as discussed in section 3.2.2.2.2.<br />

83


84 Chapter 4<br />

M/M 0<br />

1.00<br />

0.96<br />

0.92<br />

0.88<br />

0.84<br />

1.0<br />

0.9<br />

La 0.67 (Ca,Sr) 0.33 MnO 3 Magnetization<br />

0.0 0.1 0.2 0.3 0.4<br />

(T/T ) 3/2<br />

LCM Pellet (10 kOersted)<br />

LCM Film (5 kOersted)<br />

LSM Pellet (10 kOersted)<br />

LSM Film (5 kOersted)<br />

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35<br />

(T/T )<br />

C 2<br />

Figure 4-4 Magnetization of La 0.67 (Ca/Sr) 0.33 MnO 3 films and<br />

polycrystalline samples showing the T 2 dependence of the<br />

magnetization. Inset, same data as a function of T 3/2 for<br />

comparison.<br />

A T 2 dependence of the magnetization is predicted by the Moriya theory of<br />

spin fluctuation in itinerant electron ferromagnets, as shown by a calculation<br />

for the weak metallic ferromagnet Ni 3 Al [86]. A T 2 dependence has been<br />

observed in bulk SrRuO 3 (Appendices A and B) which is a metallic<br />

ferromagnet with a substantial saturation moment (i.e. p sat /p cw > 0.5) like the<br />

manganites. In the simplest model, a T 2 dependence is due to single (k-space)<br />

particle spin-flip excitations, while a T 3/2 law is expected for collective<br />

excitations (spin waves). At low temperatures, the T 3/2 law may be seen in<br />

La 0.67 Sr 0.33 MnO 3 [97] with a stiffness (D = 154 meV or Θ 3/2 = 2224 K) also<br />

observed in neutron work. This low temperature form is compared in Figure<br />

4-5 with the corresponding data of Figure 4-4 to demonstrate the limitations<br />

of this form at higher tem-peratures. The next order term in the spin wave


Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

theory predicts a T 5/2 con-tribution. Fitting the T 5/2 parameter Θ 5/2 , along with<br />

the T 3/2 term fits the data for La 0.67 Sr 0.33 MnO 3 better up to 300K than a simple<br />

T 2 fit. Some samples, such as that shown in Figure 4-5, show an increase in<br />

the magnetization at 50 K.<br />

Magnetization (emu/g/G)<br />

90<br />

85<br />

80<br />

75<br />

70<br />

65<br />

La 0.67 Sr 0.33 MnO 3<br />

M/M = 1 - (T/Θ )<br />

0 2 2<br />

Θ = 540K (free parameter)<br />

2<br />

M/M 0 = 1 - (T/Θ 3/2 ) 3/2 - (T/Θ 5/2 ) 5/2<br />

Θ 5/2 = 471K (free parameter)<br />

D = 154 meV<br />

Θ 3/2 = 2224 K<br />

60<br />

0 50 100 150 200 250 300<br />

Temperature (K)<br />

Figure 4-5 Comparison of the magnetization of<br />

La 0.67 Sr 0.33 MnO 3 with the T 3/2 term found at low temperatures,<br />

and various fits to the magnetization.<br />

4.2 Electronic Transport<br />

Both La 0.67 Sr 0.33 MnO 3 and La 0.67 Ca 0.33 MnO 3 have an abrupt drop i n<br />

resistivity below the magnetic Curie temperature in both films and bulk<br />

samples (Figure 4-6 and Figure 4-7). The magnetoresistance (Defined as R(0)-<br />

R(H)) peaks near the Curie temperature as shown in Figure 4-6 and Figure 4-7<br />

and previous experiments ([29, 104]).<br />

85


86 Chapter 4<br />

Resistivity (10 -3 Ω cm)<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Resistivity (10 -3 Ω cm)<br />

5<br />

4<br />

3<br />

2<br />

La 0.67 Sr 0.33 MnO 3 Magnetoresistance<br />

La 0.67 Sr 0.33 MnO 3 Film<br />

R (H=0)<br />

∆R/R<br />

M<br />

340 345 350 355 360 365 370 375 380<br />

Temperature(K)<br />

Pellet 0kOe<br />

Pellet 70kOe<br />

Film 0kOe<br />

Film 70kOe<br />

0<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

Figure 4-6 Magnetoresistance of La 0.67 Sr 0.33 MnO 3<br />

polycrystalline pellet and Film (LSM1). Inset, simultaneous<br />

magnetization and resistivity of the film at 20Oersted, along<br />

with the magnetoresistance [R(H = 0 kOe)-R(H = 70 kOe)].<br />

4. 2. 1 Low Temperature Resistivity<br />

Low temperature resistivity data were analyzed using a polynomial<br />

expansion in temperature T. Since the resistivity is essentially constant for<br />

temperatures less than 10K for all samples, it is clear that the resistivity data<br />

both with and without an applied field require a temperature independent<br />

(R 0 ) term. As shown previously for related compounds [29, 104, 108], R 0 is a<br />

function of magnetization in polycrystalline La 0.67 AE 0.33 MnO 3 (Figure 4-8) as<br />

discussed in section 4.2.8. In contrast, the magnetoresistance of films and<br />

crystals at 5 K is over 100 times less: it is not even measurable in the<br />

La 0.67 Ca 0.33 MnO 3 crystal where the precision is 0.5%. Due to the higher<br />

resistance of the films, the magnetoresistance at 5 K is measurable and is<br />

28<br />

26<br />

24<br />

22<br />

Magnetoresistance (%)


Resistivity (10 -3 Ω cm)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

Resistivity (10 -3 Ω cm)<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

La 0.67 Ca 0.33 MnO 3 Magnetoresistance<br />

La 0.67 Ca 0.33 MnO 3 Film<br />

Magnetization<br />

0 kOe<br />

∆R (0-70 kOe)<br />

240 245 250 255 260 265 270<br />

Temperature(K)<br />

0 kOe<br />

70 kOe<br />

0<br />

0 50 100 150 200 250 300 350 400<br />

Temperature(K)<br />

Figure 4-7 Magnetoresistance of La0.67Ca0.33MnO3 film<br />

(LCM17). Inset, simultaneous magnetization and resistivity at<br />

20Oersted, along with the magnetoresistance [R(H = 0kOe) -<br />

R( H = 70kOe)].<br />

approximately linear with field (Figure 4-8). This effect is small enough for<br />

the analysis to be carried out with R 0 independent of magnetic field.<br />

Since polycrystalline material contains a significant contribution to the<br />

resistivity from grain or domain boundaries, as shown by microwave<br />

measurements [109-111] and higher R 0 values, analysis of the resistivity data<br />

is made only for the well annealed films and crystals. Data up to 70K and<br />

100K for La 0.67 Ca 0.33 MnO 3 and La 0.67 Sr 0.33 MnO 3 , respectively, are well described<br />

with an additional (R 2 ) term proportional to T 2 (Figure 4-8). A linear term in<br />

T did not significantly improve the fit and was not used in later calculations<br />

in order to keep the number of free parameters to a minimum. The field<br />

Magnetization (relative units)<br />

87


88 Chapter 4<br />

dependence of R 2 is so small (less than 10% variation) it has not been<br />

measured.<br />

The range of validity can be extended to about 200K for La0.67Ca0.33MnO3 and 350K for La0.67Sr0.33MnO3 with the addition of a T n term where 4 < n < 5.<br />

The data was analyzed with a fit to R 0 + R 2 T 2 + R 4.5 (H)T 4.5 since a T 4.5<br />

temperature dependence has been predicted for electron-magnon scattering in<br />

the double exchange theory [112]. Evidence for a T 4.5 term (as opposed to T 4 or<br />

T 5 ) cannot be claimed from this data, only that it allows the determination of<br />

the R 0 and R 2 terms to higher accuracy in a much larger temperature range.<br />

The maximum temperature for which the data were analyzed was<br />

determined by finding the maximum temperature where the free parameters<br />

remained stable and the fit was visibly accurate.<br />

The results of this analysis are given in Table 2. As expected, R 4.5 (H)<br />

decreases as the magnetic field increases. In a field of 70 kOe, R 4.5 (H) decreases<br />

by 25-50% (Table 2).<br />

Recently other authors have analyzed the resistivity data for<br />

RE 0.67 AE 0.33 MnO 3 compounds using models different than the one presented<br />

above. The resistivity of polycrystalline La 0.67 Ca 0.33 MnO 3 has been fit to R 0 +<br />

R c T c [29] while a polynomial fit such as the one used in this work will also fit<br />

their data. The R 0 + R 2 T 2 + R 4.5 T 4.5 fit, which has the same number of free<br />

parameters, is preferable since there seems to be some universality in R 2 . The<br />

low temperature zero field resistivity of La 0.67 Ba 0.33 MnO 3 films [109] has been<br />

fit to R 0 + R 1 T + R 2 T 2 , where R 1 < 0. R 1 may not have physical meaning even<br />

though a small correction is necessary to explain the almost constant<br />

resistivity at the lowest temperatures. A dominant T 2 term has been also<br />

recently found in the resistivity of single crystals [113].


Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

4. 2. 1. 1 Temperature independent term<br />

The temperature independent term in the resistivity can be ascribed to<br />

scattering processes such as impurity, defect, grain boundary and domain wall<br />

scattering. From this value, one can make a rough estimate of the (0 K) mean<br />

free path of 26Å [56]. The temperature independent term in the<br />

polycrystalline samples is somewhat larger than that observed in films or<br />

crystals (Figure 4-6, [29]). This indicates a significant grain boundary resistance<br />

and/or a significantly restricted conduction path in the polycrystalline<br />

samples. The temperature independent terms for the various<br />

La 0.67 Ca 0.33 MnO 3 , La 0.67 Sr 0.33 MnO 3 films and crystals (Table 4-2) concur with<br />

recent measurements on other well annealed La 0.67 AE 0.33 MnO 3 [109, 114, 115]<br />

films and single crystals [116]. This limiting resistivity which is largely in-<br />

dependent of A-site atoms may be due to the inherent disorder on the A-site.<br />

Table 4-2 Magnetoresistance of Annealed Films.<br />

Sample TC (K)<br />

TMR (K)<br />

TMI (K)<br />

R 0<br />

(10 -3 Ωcm)<br />

R2 /R0 (10 -6 K -2 )<br />

R4.5 (0T)/R0 (10 -12 K -4.5 )<br />

R4.5 (7T)/R0 (10 -12 K -4.5 )<br />

Ea /kB (K)<br />

LSM1[a] 360 365 455 0.15 55 28 15<br />

LCM X 267 272 275 0.10 102 67 54 849<br />

LCM13 240 250 250 0.10 61 122 87 967<br />

LCM10 260 275 280 0.12 66 85 72 740<br />

LCM15 260 265 265 0.15 73 87 73 725<br />

LCM17 260 262 264 0.16 60 133 82 865<br />

LCM21 240 245 245 0.28 61 114 60 1116<br />

LBM[109] 310 310 330 0.34 52<br />

4.2.1.2 T 2 dependent term<br />

The T 2 term in the resistivity is also universal with respect to different<br />

samples and type of alkaline earth element (Table 2 ). Furthermore, this term<br />

is independent (within 5%) of magnetic field. The significant<br />

[a] LSM = La .67 Sr .33 MnO 3 ; LCM and LBM are the Ca and Ba analogues. X<br />

refers to float zone grown crystals.<br />

89


90 Chapter 4<br />

magnetoresistance observed in the temperature range used to fit the<br />

conductivity data is absorbed primarily in the T 4.5 term. This universality<br />

with respect to growth, composition, and magnetic field leads us to conclude<br />

that the T 2 term in the resistivity represents intrinsic behavior of<br />

La 0.67 AE 0.33 MnO 3 compounds.<br />

Resistivity (10 -3 Ω cm)<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

La 0.67 (Sr/Ca) 0.33 MnO 3 Low Temperature Resistivity<br />

LSM Film<br />

LCM Film<br />

0.1<br />

0.0 0.1 0.2 0.3<br />

(T/T )<br />

C<br />

0.4 0.5<br />

2<br />

Figure 4-8 Low temperature resistivity (in zero field) of<br />

La 0.67 (Sr/Ca) 0.33 MnO 3 films (LSM1 and LCM10). Solid lines are<br />

the fit to R 0 + R 2 T 2 + R 4.5 T 4.5 up to 250K and 200K for LSM and<br />

LCM respectively. The dashed lines show the constant and T 2<br />

terms of the best fit.<br />

Significant T 2 dependencies in resistivity [117] are often observed in<br />

metallic ferromagnets such as Fe, Co, and Ni where the coefficient [118] is 10 -11<br />

Ωcm/K 2 . The magnitude of R 2 is expected [84, 119] to scale with 1/M 0 , which<br />

explains the larger 5 x 10 -8 Ωcm/K 2 value for R 2 found in the weak itinerant


Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

electron ferromagnets Sc 3 In and ZrZn 2 . The term in La 0.67 AE 0.33 MnO 3 (Table<br />

4-2) is about 1000 times larger than that in Fe, Co and Ni, which have<br />

comparable M 0 's. Furthermore, the value of R 2 is expected to decrease in a<br />

magnetic field (proportional to H -1/3 ) due to the suppression of spin<br />

fluctuations; however, the field dependence of R 2 in the manganites is too<br />

small for us to detect. Thus the theory of spin fluctuations does not<br />

completely explain the observed T 2 dependence of the resistivity.<br />

General electron-electron scattering mechanisms within the Fermi liquid<br />

model give a T 2 dependence of the resistivity which is not necessarily field<br />

dependent. A value of R 2 as large as that seen in this work has been found in<br />

the nonmagnetic semimetal TiS 2 [120].<br />

4. 2. 1. 3 Relationship to magnetism<br />

A quite interesting correlation [114, 115, 121, 122] has been found between<br />

the magnetization and the resistivity, namely ρ = R e exp(-M(T, H)/M r ) where<br />

R e and M r are fitting parameters. This is clearly preferable to a polynomial fit<br />

in the critical region near T C . If, at low temperatures, M = M 0 (1-M 2 T 2 ) where<br />

M 2 is largely independent of field, then this expression predicts the resistivity<br />

will vary as R 2 T 2 + R 4 T 4 . Thus a three parameter fit to ρ =<br />

R e exp(-M(T, H)/M r ) is essentially equivalent to R 0 + R 2 T 2 + R 4 T 4 and<br />

therefore fits our data very well. However, this equivalency breaks down<br />

when one considers the field dependence: Since both R e and M r are<br />

independent of field, the field dependence of ρ comes from the field<br />

dependence of M 2 which would affect R 2 and R 4 equally. In summary, the<br />

exponential fit combined with a T 2 dependence of the magnetization uses<br />

four free parameters: R 0 , R e , M r and M 2 (7T) which does not fit the low<br />

temperature data as well as the four free parameters R 0 , R 2 , R 4.5 (0T), R 4.5 (7T)<br />

used in this work. This is discussed further in chapter 7.<br />

91


92 Chapter 4<br />

4. 2. 2 High Temperature resistivity<br />

Above T C , The zero field resistivity (Figure 4-9) of La 0.67 Ca 0.33 MnO 3 fit best<br />

to R h Texp(E a /k b T) predicted by small polaron hopping conductivity [123], as<br />

opposed to that often used for a semiconductor, ln(ρ) ∝ 1/T (Figure 9). Other<br />

(low temperature) forms predicted by variable range hopping [124] or band<br />

semiconductor type conductivity were also fit with two parameters to the data<br />

with less accurate and less consistent results. In contrast to the Ca and Ba<br />

analogs [29, 104] the La 0.67 Sr 0.33 MnO 3 resistivity (Figure 4-10) does not reach a<br />

maximum until about 100K above T C [16].<br />

ln(ρ(10 -3 Ωcm)/T(K))<br />

-3.0<br />

-3.5<br />

-4.0<br />

-4.5<br />

-5.0<br />

-5.5<br />

ln(ρ(10 -3 Ωcm))<br />

2.4<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

750K<br />

La 0.67 Ca 0.33 MnO 3<br />

Crystal<br />

Film<br />

1.0 1.5 2.0 2.5 3.0 3.5<br />

1000/T (1/K)<br />

750K<br />

a<br />

Crystal<br />

Film<br />

-6.0<br />

0.5 1.0 1.5 2.0<br />

1000/T (1/K)<br />

2.5 3.0 3.5<br />

mW/mg<br />

0.09<br />

0.08<br />

0.07<br />

b<br />

730 740 750 760 770 780<br />

Temperature (K)<br />

Figure 4-9 High temperature resistivity (warming and<br />

cooling) of La 0.67 Ca 0.33 MnO 3 film (LCM17) and crystal in zero<br />

field. Inset a, same data with different abscissa to compare<br />

small polaron and semiconductor models. Inset b, DSC trace<br />

of polycrystalline La 0.67 Ca 0.33 MnO 3 showing the heat of the<br />

high temperature structural transition.


Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

In the theory of small polaron conductivity, E a is the hopping energy. The<br />

prefactor R h ≈ k b /(e 2 a 2 p(1-x)ν) where k b is Boltzmann’s constant, e is the charge<br />

of the hole, a is the distance between hopping sites, p is the carrier density, x is<br />

the doping fraction (1/3 in this case) and ν is the optical phonon or attempt<br />

frequency. A least squares fit of the high temperature data gives a hopping<br />

activation energy of 865K and attempt frequency ν = 6·10 13 /s.<br />

Resistivity (10 -3 Ωcm)<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

T C<br />

ln( ρ(10 -3 Ωcm)/T(K))<br />

-4.6<br />

-4.8<br />

-5.0<br />

-5.2<br />

-5.4<br />

-5.6<br />

-5.8<br />

La 0.67 Sr 0.33 MnO 3 Film<br />

1.0 1.5 2.0 2.5 3.0 3.5<br />

1000/T (1/K)<br />

1.5<br />

300 400 500 600 700 800 900 1000 1100<br />

Temperature (K)<br />

Figure 4-10 High temperature resistivity (warming and<br />

cooling) of La 0.67 Sr 0.33 MnO 3 film (LSM1) in zero field. Inset,<br />

same data displayed as in Figure 4-9.<br />

4. 2. 3 Transport Near T C<br />

The resistance of La 0.67 Ca 0.33 MnO 3 reaches a maximum near the Curie<br />

temperature whereas that of La 0.67 Sr 0.33 MnO 3 maximizes about 100K above T C .<br />

This is best seen from simultaneous magnetization and resistance<br />

93


94 Chapter 4<br />

measurements (Figure 4-6 and Figure 4-7), where there is no error in relative<br />

temperature. Both La 0.67 Ca 0.33 MnO 3 and La 0.67 Sr 0.33 MnO 3 have T MR ≈ T C<br />

within a few Kelvin. For La 0.67 Ca 0.33 MnO 3 T MI is approximately equal to T C ≈<br />

T MR . La 0.67 Sr 0.33 MnO 3 however shows a much more gradual transition to a<br />

hopping conductivity-like transport with T MI (approximately 455K) well above<br />

T C = 360K.<br />

At T C a maximum in dρ(H=0)/dT is expected for a metallic ferromagnet<br />

[125]. This is observed for both compounds studied within experimental<br />

uncertainty. The added resistance at a ferromagnetic Curie temperature is<br />

due to electron scattering off thermally disordered spins and, particularly in<br />

the case of the manganites, polaron formation. Since a magnetic field can<br />

easily suppress spin fluctuations in the critical region, the resistance<br />

associated with magnetic disorder will be reduced in a magnetic field. In a<br />

good metal such as Fe or SrRuO 3 [126] this normally is a small effect of about a<br />

few percent. It has been shown theoretically that this effect is much larger for<br />

a semimetal (or semiconductor) at T C , particularly within the double<br />

exchange model [8, 125, 127, 128]; however, it has been recently been pointed<br />

out that double exchange alone can not account for the large<br />

magnetoresistance [10, 11]. Nevertheless, such an explanation has been used<br />

to explain the magneto-transport properties of semiconducting magnetic<br />

chalcogenides [129] such as EuO [130], Gd 2 S 3 [131], and various spinels [129, 132-<br />

135] where the resistance may drop by several orders of magnitude at T C .<br />

4. 2. 4 Hall effect<br />

The Hall effect data at 5 K (Figure 4-11) were analyzed according to the<br />

equation R = R 0 + H·R H + |H|·R MR , where R H is the Hall resistance and R MR is<br />

the magnetoresistance. The contribution due to the anomalous Hall effect<br />

was not detected at this temperature. From the simple single band interpreta-


Resistance (Ω)<br />

2.88<br />

2.86<br />

2.84<br />

2.82<br />

2.80<br />

Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

La 0.67 (Ca/Sr) 0.33 MnO 3 Hall Effect at 5K<br />

La/Ca Film<br />

La/Sr Film<br />

Hall Effect<br />

-60 -40 -20 0 20 40 60<br />

Field (kOe)<br />

tion of the Hall effect [56], our measurements at 5 K show hole conductivity<br />

with concentrations of the expected order of magnitude (Table 1).<br />

Because the Hall effect is so small and conduction appears to proceed via<br />

small polaron hopping, it is concluded that the temperature dependence of<br />

the resistivity is due to the temperature dependence of the mobility while the<br />

carrier concentration remains relatively constant.<br />

Since LaMnO 3 is a high spin d 4 insulator, the e g orbitals must be split,<br />

possibly by the Jahn-Teller effect. Thus doping with an alkaline earth<br />

element on the La site, should put mobile holes in the lower e g state. The<br />

magnitude of the carrier concentration calculated from the Hall effect is<br />

2.02<br />

2.01<br />

2.00<br />

1.99<br />

1.98<br />

Figure 4-11 Resistance as a function of field<br />

La 0.67 (Ca/Sr) 0.33 MnO 3 films (LSM1 and LCM19) in the Hall<br />

effect configuration at 5 K. The Hall effect is calculated from<br />

the slope of the line indicated (see text).<br />

95


96 Chapter 4<br />

somewhat too large compared with the above model and therefore such a<br />

simple interpretation is clearly inadequate.<br />

4. 2. 5 Crystallographic Phase change<br />

At 750K, a nearly discontinuous jump as well as a significant change in<br />

slope is observed in the resistivity. The jump is much more noticeable in the<br />

crystals indicating a structural transition which is suppressed to some extent<br />

by the epitaxy in the thin films. This transition has been confirmed by<br />

differential scanning calorimetry (Figure 4-9). An orthorhombic to rhombo-<br />

hedral transition is seen in at low temperatures in La 0.83 Sr 0.17 MnO 3 [126].<br />

A change in slope is only noticeable in the ln(ρ) vs. 1/T plot but not in the<br />

ln(ρ/T) vs. 1/T plot. This indicates that the change in dρ/dT can be<br />

interpreted within the small polaron hopping conductivity model [62]. Below<br />

750K thermally activated hopping of the charge carriers dominates the<br />

conductivity, while above 750K scattering of these carriers begins to dominate.<br />

4. 2. 6 Small Polarons<br />

The presence of small polarons implies the existence of local lattice<br />

distortions which localize the charge carriers. A Jahn-Teller distortion about<br />

the Mn 3+ ion is expected from its d 4 configuration in an octahedral<br />

environment. It has been postulated that such a distortion is necessary to<br />

explain the magnetoresistivity of these compounds [10, 11].<br />

Above TC the resistivity of La0.67Sr0.33MnO3 may be explained by a<br />

crossover between two types of polaron conduction [116]. Just above TC the<br />

polarons are heavy free carriers scattered by phonons, which gives a positive<br />

dρ/dT. Quantitatively, this is expected to add a term to the conductivity<br />

proportional to exp(-T/Θ D )/T [62]. At T MI , which should be about half the<br />

Debye temperature (Θ D ), the phonon scattering becomes so strong that the<br />

mean free path is about one lattice spacing, and therefore localizes the


Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

carriers. Above T MI the polarons conduct via thermally activated hopping,<br />

which has a negative dρ/dT. Although this qualitatively explains the data,<br />

the quantitative predication of the temperature dependence does not fit the<br />

data very well.<br />

4. 2. 7 Colossal Magnetoresistance<br />

In general, manganite films reported in the literature which exhibit CMR<br />

as shown in Figure 4-12 [21, 23, 26, 127, 128]: 1) do not have square hysteresis<br />

loops at low temperatures, leading some to believe superparamagnetic<br />

behavior exists in those materials; 2) have ill-defined T C 's which are much<br />

less than those reported for the bulk materials with the same nominal<br />

composition; 3) have T MI well below T C , with T MR slightly less than T MI ; and<br />

4) have high resistivities in the metallic state, at low temperatures.<br />

Figure 4-12 Colossal magnetoresistive La 0.67 Ca 0.33 MnO 3 film<br />

from [21, 127].<br />

The annealed films in this study (which do not exhibit CMR) are much<br />

like bulk polycrystalline or single crystal material: They are clearly itinerant<br />

97


98 Chapter 4<br />

Magnetization<br />

8<br />

6<br />

4<br />

2<br />

La 0.67 Sr 0.33 MnO 3 Polycrystalline Pellet<br />

Magnetization<br />

0<br />

0 2 4 6 8 10<br />

Field (kOe)<br />

electron ferromagnets with reasonably square hysteresis loops, T 2 dependence<br />

of the saturation magnetization and sharp T C 's. They have T C and T MR all<br />

approximately equal and along with T MI are comparable to those of the bulk<br />

polycrystalline material. Finally, the films have low resistivities like those<br />

observed in single crystals [104, 109, 114, 129-131]. It is therefore concluded<br />

that these films are displaying properties inherent to La 0.67 (Sr/Ca) 0.33 MnO 3 and<br />

less influenced by microstructure, strain, and/or compositional variations<br />

than films with suppressed T C and CMR.<br />

Resistivity (10 -3 Ω cm)<br />

0.1248<br />

0.1246<br />

0.1244<br />

0.1242<br />

0.1240<br />

0.1238<br />

Resistivity<br />

La 0.67 Ca 0.33 MnO 3 Film<br />

0.1236<br />

0 10 20 30 40 50 60 70<br />

Field (kOe)<br />

If one assumes that the resistivity behaves as Texp(E a /k B T) above T MI and at<br />

least R 2 T 2 in an infinite field below T C , then the maximum magnetoresistance<br />

0.68<br />

0.64<br />

0.60<br />

0.56<br />

0.52<br />

0.48<br />

Figure 4-13 Simultaneous magnetization and resistivity in a<br />

magnetic field of La 0.67 Sr 0.33 MnO 3 polycrystalline pellet at 5 K.<br />

Data for both increasing and decreasing field are shown. Inset,<br />

Magnetoresistance of La 0.67 Ca 0.33 MnO 3 film (LCM10) at 5 K.<br />

Resistivity (10 -3 Ωcm)


Intrinsic Properties of La 0.67 Ca 0.33 MnO 3<br />

(∆R/R(H) at T = T MR = T MI ) follows exp(E a /k B T MI )/(T MI ). This predicts<br />

"colossal" magneto resistance only at low temperatures, as a consequence of<br />

the depressed T MI . Such a strong dependence of the magnetoresistance on a<br />

film's particular T MI has been documented (Appendices A and B) from the<br />

various values reported in the literature.<br />

4.2.8 Domain Boundary Magnetoresistance<br />

The negative magnetoresistance associated with the domain orientation is<br />

only observed in polycrystalline samples [28, 96]. If this is due to domain wall<br />

scattering, then films and crystals with fewer domain boundaries would be<br />

expected to show a much smaller effect. It has been concluded that spin-<br />

polarized inter-grain tunneling is responsible for this effect [132].<br />

4. 2. 9 Low temperature magnetoresistance<br />

The low temperature negative magnetoresistance linear in H is very small<br />

in the films but somewhat larger in the polycrystalline samples (Figure 4-13,<br />

and [28, 96]). This negative magnetoresistance may be due to the M 2<br />

dependence of the resistivity described in section 7.2.4 [129, 130] combined<br />

with a small forced magnetization. Classic magnetoresistance of metals is<br />

generally positive but is often negative in ferromagnets.<br />

4.3 Conclusion<br />

Well annealed MOCVD thin films show properties of bulk<br />

La 0.67 Ca 0.33 MnO 3 and La 0.67 Sr 0.33 MnO 3 . The limiting low temperature<br />

resistivity of 0.2 mΩcm, which gives a mean free path of roughly 20Å, is<br />

independent of alkaline earth element and only slightly dependent on field<br />

for single crystal material (bulk or film). The manganites show a significant<br />

T 2 dependence of the resistivity which is also independent of alkaline earth<br />

element and magnetic field. The maximum in the magnetoresistance occurs<br />

at the Curie temperature, which is not necessarily where dρ/dT changes sign,<br />

but where dρ/dT is a maximum. A large magnetoresistance can occur well<br />

99


100 Chapter 4<br />

above room temperature, but the effect decreases significantly with<br />

temperature.<br />

Three regions of magnetoresistance are identified in these materials. The<br />

largest effect is likely due to the suppression of magnetic critical scattering<br />

near the Curie temperature. There is also a negative magnetoresistance<br />

associated with the net magnetization of polycrystalline samples but not seen<br />

in single crystal materials. Finally there is a small negative<br />

magnetoresistance linear in field even at low temperatures.<br />

The compounds La 0.67 Ca 0.33 MnO 3 and La 0.67 Sr 0.33 MnO 3 are metallic<br />

ferromagnets with large saturation moments. The saturation magnetization<br />

is found to decrease proportional to T 2 .<br />

The high temperature resistivity of La 0.67 Ca 0.33 MnO 3 clearly follows the law<br />

predicted by small polaron hopping conductivity both in the thermally<br />

activated regime and at higher temperatures where scattering becomes<br />

important. There appears to be a structural transition at about 750K.<br />

The resistivity of La 0.67 Sr 0.33 MnO 3 above the Curie temperature shows a<br />

crossover from a metallic to a hopping regime at higher temperatures.<br />

In conclusion, the CMR effects in materials of the same compositions as<br />

those studied here but with much lower transition temperatures, are not<br />

intrinsic to the thermodynamically stable phase. Local inhomogeneities or<br />

noncrystallinity in not fully annealed films may suppress the magnetic and<br />

metal-insulator transitions. This presumably causes the observed<br />

superparamagnetic type magnetism and conduction via percolation.


5. Local Structure, Transport and Rare Earth Magnetism in<br />

the Ferrimagnetic Perovskite Gd0.67Ca0.33MnO3 In order to examine the nature of the metal-insulator transition in the<br />

CMR manganites, the distorted perovskite Gd 0.67 Ca 0.33 MnO 3 related to<br />

La 0.67 Ca 0.33 MnO 3 was studied. The composition x = 1/3 was chosen because<br />

this doping concentration should maximize the double exchange effect as<br />

seen in the lanthanum compounds [29]. Due to the small size of the Gd 3+<br />

ion, Gd 0.67 Ca 0.33 MnO 3 should be in the region of the x = 1/3 phase diagram<br />

where there is no transition to a metallic state, but to a ferromagnetic<br />

insulating state around 50K [100]. The ferromagnetic nature of this low<br />

temperature state has recently been questioned due to evidence of a spin glass<br />

behavior with no long range ferromagnetic order in (Tb x La 1-x ) 0.67 Ca 0.33 MnO 3<br />

[136]. The x-ray-absorption fine-structure (XAFS) technique can detect small<br />

variations of the average local environment about a particular atomic species,<br />

making it ideal for studying subtle structural phase transitions associated with<br />

any metal-insulator or ferromagnetic transitions. The Mn K-edge (6.54 keV)<br />

and Gd L III -edge (7.25 keV) are well enough separated that there is little<br />

interference of the Mn and Gd XAFS.<br />

In this study, it is found that the rare earth moments in Gd 0.67 Ca 0.33 MnO 3<br />

order antiparallel to the manganese giving rise to ferrimagnetism.<br />

Furthermore, not only does this material remain insulating when it becomes<br />

magnetically ordered, but no structural difference exists between the<br />

ferrimagnetic and paramagnetic state. This is consistent with models that<br />

require a structural change, as well as the ferromagnetic ordering of the<br />

manganese atoms, to explain the metal-insulator transition and the large<br />

magnetoresistance that accompanies it. Most of the work presented in this<br />

chapter has been previously published [137].<br />

101


102 Chapter 5<br />

5.1 Ferrimagnetism<br />

The polycrystalline pellets, crystals and thin films all display a<br />

ferromagnetic transition and a compensation temperature (Figure 5-1), which<br />

is characteristic of ferrimagnets. Both bulk materials, polycrystalline pellet<br />

and float zone crystal, have very similar magnetic properties while the film is<br />

slightly different.<br />

Gd 0.67 Ca 0.33 MnO 3 displays the properties of a ferrimagnet, much like the<br />

garnet ferrite Gd 3 Fe 5 O 3 [138]. At about 60K, the transition metal spins order,<br />

dominating the magnetization. As the temperature is lowered, the opposing<br />

magnetization of the weaker coupled rare earth ions increases, since they<br />

behave almost as though they were free paramagnetic spins reacting to an<br />

internal field (section 5.1.3, Figure 3-19).<br />

At the compensation temperature, T Comp , the rare earth and transition<br />

metal sublattice magnetizations exactly cancel (in zero field). Below T Comp , the<br />

magnetization of the rare earth ions, which have larger spins, dominates.<br />

This is best seen in zero field remnant data which shows the magnetization<br />

changing sign at T Comp .<br />

In the presence of magnetic fields greater than the coercive field, the<br />

stronger sublattice flips to align parallel with the field, giving the minimum<br />

in M at T Comp as seen the 5 kOe data.


Magnetization (µ B /mole)<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

-0.2<br />

Local Structure and Magnetism in Gd 0.67 Ca 0.33 MnO 3<br />

Gd 0.67 Ca 0.33 MnO 3 Magnetization<br />

5kOe<br />

Crystal<br />

Pellet<br />

Film<br />

Remnant<br />

× 25<br />

0 10 20 30 40 50 60 70 80 90<br />

Temperature (K)<br />

Figure 5-1 Low temperature magnetization of<br />

Gd 0.67 Ca 0.33 MnO 3 measured in a 5 kOe field and zero field after<br />

cooling in a large field (remnant).<br />

5. 1. 1 Low temperature moment<br />

The saturation magnetization per mole of Gd 0.67 Ca 0.33 MnO 3 expected from<br />

the Gd 3+ S = 7/2, l = 0 is µ=(0.67 × 2 × 7/2) µ B = 4.67µ B , while high spin<br />

manganese gives spin only (orbital contribution quenched) µ = gSµ B , g = 2 so µ<br />

= 2µ B [0.67 × 2 (from Mn 3+ ) + 0.33 × 3/2 (from Mn 4+ )] = 3.67µ B . The simple two<br />

sublattice (Gd and Mn) ferrimagnetism described above predicts a zero<br />

temperature saturation magnetization of 4.67 µ B - 3.67 µ B = 1.00 µ B (if there is<br />

any canting of the sublattices such as that in Pr 0.67 Ca 0.33 MnO 3 [139], a smaller<br />

moment is predicted). This is roughly what is observed. The mean field<br />

103


104 Chapter 5<br />

calculation predicts a constant moment for T < 5 K, while the data show an<br />

almost linear change in M for 2 K < T < 5 K. Also, the peak in the<br />

magnetization, observed at about 50 K in the 5 kOe data is less than the 1.9µ B<br />

predicted in the mean field calculation.<br />

5. 1. 2 High temperature susceptibility<br />

Susceptibility data (Figure 5-2) above 200 K fit to the Curie-Weiss law:<br />

2 2 2<br />

χ = µ eff /(8(T-Θ)) with Θ = 34 K and µeff = 71 µB . The difference in the<br />

susceptibility of the crystal compared to that of the pellet at high temperatures<br />

2<br />

may not be significant. For quantum mechanical spins µ eff per mole is<br />

expected to be 63 µ B 2 : 2 2 µB 2 (0.67 × 4/2 × 6/2) = 16 µB 2 from Mn 3+ ; 2 2 µB 2 (0.33 ×<br />

3/2 × 5/2) = 5 µ B 2 from Mn 4+ ; 2 2 µB 2 (0.67 × 7/2 × 9/2) = 42 µB 2 from Gd 3+ . This<br />

extra moment, which is also observed in La 0.67 Ca 0.33 MnO 3 [103] section 4.1,<br />

may be due to the orbital contribution in the manganese ions. Since the two<br />

sublattices have different coupling energies, the susceptibility above T C will<br />

not obey a simple Curie-Weiss law even in the mean field approximation<br />

(Figure 5-5). The simple mean field model of the above ferrimagnetism<br />

requires T C /Θ > 3, as described in section 3.2.2.4.1. This is clearly not the case<br />

for the measured parameters. Alternatively, the addition of a Pauli-like<br />

2<br />

susceptibility χ0 to the fit, χ = χ0 + µ eff /(8(T-Θ)), will reduce the contribution<br />

from µ eff . For example, an equally good fit to the data can be found with<br />

2 2<br />

µ eff = 63 µB , χ0 = 7.7 × 10 -6 emu G -1 g -1 , and Θ = 50 K.


1/χ mol (Gauss·mol/emu)<br />

Local Structure and Magnetism in Gd 0.67 Ca 0.33 MnO 3<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Gd 0.67 Ca 0.33 MnO 3 Susceptibility<br />

Pellet<br />

Crystal<br />

5 kOe<br />

0<br />

0 50 100 150 200 250 300 350 400<br />

Temperature (K)<br />

Figure 5-2 Inverse magnetic susceptibility of bulk<br />

Gd0.67Ca0.33MnO3 . Solid line is the high temperature fit to χ =<br />

2<br />

µ eff /(8(T-Θ)) described in the text.<br />

5. 1. 3 Low temperature susceptibility<br />

Between T Comp and T C , the magnetic susceptibility essentially that of the<br />

paramagnetic-like rare earth moments, showing an approximate<br />

χ ∝ 1/(T + T Comp ). Below T Comp the Gd 3+ moments effectively order antiferro-<br />

magnetically with respect to the manganese moments, and so the<br />

susceptibility decreases below T Comp .<br />

105


106 Chapter 5<br />

χ mol (µ B /mol/T)<br />

0.40<br />

0.36<br />

0.32<br />

0.28<br />

0.24<br />

0.20<br />

High Field Susceptibility of Gd 0.67 Ca 0.33 MnO 3<br />

Magnetization (µ B /mole)<br />

0 10 20 30 40 50 60 70 80<br />

Temperature (K)<br />

3<br />

2<br />

1<br />

0<br />

5K<br />

0 10 20 30 40 50 60 70<br />

Magnetic Field (kOe)<br />

Figure 5-3 Low temperature and high-field magnetic<br />

susceptibility, χ = (M(60 kOe)-M(40 kOe))/20kOe, of<br />

Gd 0.67 Ca 0.33 MnO 3 crystal. Inset, hysteresis loop at 5 K.<br />

The hysteresis loop at 5 K (Figure 5-3) shows an additional high field<br />

paramagnetic response not expected for a mean field ferrimagnet. It is not<br />

uncommon, however, to observe such a field-induced increase of the<br />

magnetization in canted ferromagnetic or ferrimagnetic systems [140, 141],<br />

and is used as evidence of the ferrimagnetic sublattices canting away from<br />

antiparallel [101]. The magnetic field affects the canting angle, which results<br />

in the magnetization increasing nearly linearly with field even at T = 0 K<br />

[140]. If this large susceptibility were due to a second phase with free<br />

paramagnetic spins, then this phase would have to contain at least half of the<br />

available spins and have a 1/T temperature dependence for all T. This is not


Magnetization 2 (µ B /mole) 2<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

Local Structure and Magnetism in Gd 0.67 Ca 0.33 MnO 3<br />

Gd 0.67 Ca 0.33 MnO 3 Arrott Plot<br />

35K<br />

40K<br />

45K<br />

50K<br />

55K<br />

60K<br />

70K<br />

0.00<br />

0 2000 4000 6000 8000 10000<br />

H/M (Oe/µ )<br />

B<br />

Figure 5-4 Arrott plot of polycrystalline Gd 0.67 Ca 0.33 MnO 3<br />

pellet.<br />

seen in Figure 5-3. Instead, the high field susceptibility maximizes near T Comp<br />

which is typical of an antiferromagnet with T N ≈ T Comp . This is predicted in the<br />

mean field calculation for a ferrimagnet described in section 3.2.2.4.1.<br />

5. 1. 4 Near T C magnetism<br />

The experimental values of T C are summarized in Table 5-1. An example<br />

Table 5-1 Transition Temperatures for Gd 0.67 Ca 0.33 MnO 3 .<br />

T Comp Arrott T C Remnant T C Inflection T C<br />

Polycrystalline Pellet 17K 52K 60K 80K<br />

Single Crystal 17K 56K 62K 78K<br />

Thin Film 13K 70K 81K 88K<br />

107


108 Chapter 5<br />

of the nonlinear M 2 vs. H/M Arrott plot is given in Figure 5-4.<br />

5. 1. 5 Magnetism model<br />

To explain the magnetism, a mean field model was used assuming a<br />

ferrimagnetic ground state. Other possible explanations for the magnetism in<br />

Gd 0.67 Ca 0.33 MnO 3 are briefly discussed<br />

5 kOe<br />

Gd 0.67 Ca 0.33 MnO 3 Mean Field Calculation<br />

Remnant<br />

Gd<br />

Mn<br />

20 40 60 80 100 -5<br />

Temperature (K)<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

Magnetization (µ B /mol)<br />

5 kOe<br />

0 50 100 150 200 250<br />

0<br />

300<br />

Temperature (K)<br />

Figure 5-5 Magnetization and inverse magnetic susceptibility<br />

calculated for Gd 0.67 Ca 0.33 MnO 3 using the simplified mean field<br />

theory described in the text and T C = 83 K, T Comp = 17 K. The<br />

contribution to the magnetization of each sublattice is shown<br />

in dashed lines.<br />

5. 1. 5. 1 Mean Field Model<br />

Due to the stronger 3d exchange interaction compared to that of the 4f, the<br />

Mn-Mn coupling is expected to be stronger than the Mn-Gd, and the Gd-Gd<br />

interaction to be negligible. Since the observed Curie temperature is about<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

1/χ (mol G/emu)


Local Structure and Magnetism in Gd 0.67 Ca 0.33 MnO 3<br />

that expected from the trend found in the manganese ferromagnetic ordering<br />

temperature as a function of the average of the R and A atom sizes in<br />

R 1-x A x MnO 3 [100], it is assumed for now that the Mn-Mn interaction in<br />

Gd 0.67 Ca 0.33 MnO 3 is ferromagnetic. The ferrimagnetism must then arise from<br />

a weaker, antiferromagnetic Mn-Gd interaction. Thus, unlike the ferrites,<br />

where the dominant interaction is antiferromagnetic between the two<br />

transition metal sublattices, the strongest magnetic coupling in<br />

Gd 0.67 Ca 0.33 MnO 3 is ferromagnetic, much like that found in the rare earth<br />

cobaltites, e. g. GdCo 5 [142].<br />

The above description of the magnetization can be bolstered using a<br />

simple molecular or mean field model described in section 3.2.2.4. Assuming<br />

that all atoms on each sublattice (Mn or Gd) feel the same mean field, there<br />

are only two unknown mean field coefficients due to the Mn-Mn and Mn-Gd<br />

interactions. These two coefficients are easily determined since they are<br />

nearly proportional to T C and T Comp , respectively. This molecular field model<br />

(Figure 5-5) reproduces the qualitative features described above.<br />

The calculated magnetization and inverse susceptibility for the molecular<br />

field model described below is shown in Figure 5-5.<br />

This model is limited by the obvious simplifications: not only are there<br />

two distinct types of Mn atoms (Mn 3+ and Mn 4+ ), but there are numerous<br />

possible configurations for the number and type of near neighbors due to the<br />

randomness on the Mn and Gd/Ca sites. The discrepancy among the various<br />

values of T C determined by the three methods mentioned is an example of<br />

non-mean-field ferromagnetic behavior. The discrepancies in the salient<br />

features discussed above may result from the simplifications of the model,<br />

and the assumptions behind it.<br />

5. 1. 5. 2 Canted antiferromagnetism<br />

The evidence concerning the magnetic structure of the manganese in the<br />

ferromagnetic insulator phase points to a canted antiferromagnetic structure<br />

109


110 Chapter 5<br />

rather than a simple ferromagnet. The magnetic structure of Pr 0.67 Ca 0.33 MnO 3<br />

which is on the boundary between ferromagnetic-insulator and ferro-<br />

magnetic-metal low temperature phases [100], has been extensively studied.<br />

Neutron diffraction shows Pr 0.67 Ca 0.33 MnO 3 to have a charge-ordered, canted-<br />

antiferromagnetic structure in the low temperature phase with ferromagnetic<br />

moment about half of the total expected moment [139, 143]. A compound<br />

clearly in the ferromagnetic-insulator part of the tolerance-factor phase-<br />

diagram (Figure 3-17), much closer to Gd 0.67 Ca 0.33 MnO 3 than Pr 0.67 Ca 0.33 MnO 3<br />

is La 0.2 Y 0.5 Ca 0.3 MnO 3 . This compound appears to be ferromagnetic with about<br />

half the expected ferromagnetic moment [100], which could be explained by a<br />

canted antiferromagnetic state similar to that found in Pr 0.67 Ca 0.33 MnO 3 . Thus,<br />

one should expect the ferromagnetic moment of the Mn ions in<br />

Gd 0.67 Ca 0.33 MnO 3 to be half that predicted in the above molecular field model.<br />

If the Gd moments are also canted to the same degree, then the predicted<br />

magnetization will be qualitatively the same for the uncanted case<br />

(Figure 5-5), only half the magnitude; namely, there will still be a<br />

compensation temperature.<br />

5. 1. 5. 3 Spin glass magnetism<br />

Atomic disorder such as that found on both the Gd and Mn site in<br />

Gd 0.67 Ca 0.33 MnO 3 can result in a spin-glass, and indeed a spin-glass insulating<br />

low-temperature phase was found in (Tb x La 1-x ) 0.67 Ca 0.33 MnO 3 in place of the<br />

ferromagnetic insulator state. The compound Tb 0.67 Ca 0.33 MnO 3 should be in<br />

the same region of the tolerance-factor phase-diagram as Gd 0.67 Ca 0.33 MnO 3 and<br />

La 0.2 Y 0.5 Ca 0.3 MnO 3 . The Tb compound is apparently a pure spin-glass with no<br />

long-range magnetic order [136]. A full investigation of the existence or non-<br />

existence of spin-glass behavior in Gd 0.67 Ca 0.33 MnO 3 is beyond the scope of this<br />

work. However, it is clear that Gd 0.67 Ca 0.33 MnO 3 has long range magnetic<br />

order, which eliminates the possibility of a spin glass. The distinction is most<br />

obvious in the Arrott plot (Figure 4), which shows that Gd 0.67 Ca 0.33 MnO 3 has a


ln(ρ/T (Ω cm/K))<br />

Local Structure and Magnetism in Gd 0.67 Ca 0.33 MnO 3<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

ln(ρ/Ω cm)<br />

Gd 0.67 Ca 0.33 MnO 3 High Temp. Resistivity<br />

4<br />

3<br />

2<br />

1<br />

1.0 1.5 2.0 2.5 3.0<br />

1000/T (1000/K)<br />

1 2 3 4 5 6 7<br />

1000/T (1000/K)<br />

non-zero magnetization when H = 0 below T C , while<br />

(Tb 0.33 La 0.67 ) 0.67 Ca 0.33 MnO 3 does not [136]. The other striking feature of<br />

Gd 0.67 Ca 0.33 MnO 3 is the existence of a compensation point, which to our<br />

knowledge cannot be explained by a spin-glass magnetic system.<br />

5.1.5.4 Related Compounds<br />

Since Gd 3+ has the largest spin of the rare earth elements, it should have<br />

the highest ordering temperature. If the relevant exchange interaction is due<br />

to spin-spin coupling of the rare earth and manganese atoms, then the<br />

strength of this interaction will be proportional to the spin S of the rare earth<br />

a<br />

ln(ρ/Ω cm)<br />

4<br />

3<br />

2<br />

1<br />

b<br />

0.18 0.20 0.22<br />

1/T 1/4 (K -1/4 )<br />

Figure 5-6 High temperature resistivity during heating and<br />

cooling a Gd 0.67 Ca 0.33 MnO 3 film, ln(ρ/T) vs. 1/T. Inset a,<br />

comparison with ln(ρ) vs. 1/T. Inset b, comparison with ln(ρ)<br />

vs. 1/T 1/4 .<br />

111


112 Chapter 5<br />

ln(ρ /Ω cm)<br />

24<br />

22<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

Gd 0.67 Ca 0.33 MnO 3 Low Temperature Resistivity<br />

ln(ρ/T (Ω cm/K))<br />

20<br />

15<br />

10<br />

ion. From this relation the ordering temperature of the rare earth moments<br />

in related compounds is estimated to be about 7 K for Nd 0.67 Ca 0.33 MnO 3 and<br />

5 K for Pr 0.67 Ca 0.33 MnO 3 . This should result in a decrease in M at low T. This<br />

appears at temperatures lower than these estimates, if at all [23, 144]; perhaps<br />

indicating an interaction strength proportional to S 2 , the square of the rare<br />

earth spin, or simply that these moments do not order antiparallel to the<br />

manganese moments [145].<br />

5<br />

10 12 14 16 18 20 22<br />

1000/T (1/K)<br />

65 K ≈ Tc<br />

0.32 0.33 0.34 0.35 0.36 0.37 0.38<br />

a<br />

0.0<br />

-0.2<br />

-0.4<br />

-0.6<br />

1/T 1/4 (K -1/4 )<br />

300K MR (%)<br />

b<br />

300K<br />

200K<br />

200K MR (%)<br />

-1<br />

-2<br />

-0.8<br />

-3<br />

-60 -40 -20 0 20 40 60<br />

Field (kOe)<br />

Figure 5-7 Low temperature resistivity of Gd 0.67 Ca 0.33 MnO 3<br />

crystal, ln(ρ) vs. 1/T 1/4 . Inset a, comparison with ln(ρ/T) vs.<br />

1/T. Solid lines show linear best fit to the data shown. Inset b,<br />

magnetoresistance of a film at 200 K and 300 K; solid line is<br />

the quadratic fit.<br />

0


Local Structure and Magnetism in Gd 0.67 Ca 0.33 MnO 3<br />

5.2 Electronic Transport<br />

The resistivity can be well approximated by R h Texp(E a /k B T) (Figure 5-6),<br />

predicted by small polaron hopping conductivity [68, 146]. Below about 150 K<br />

(Figure 5-7), the resistivity is somewhat better described by R 0 exp(T 0 /T) 1/4 ,<br />

which is commonly attributed to variable range hopping [123]. The<br />

quantitative aspects of the fits is discussed below. The thin film and crystal<br />

samples show no transition to a metallic state down to 5 K.<br />

5.2.1 Magnetoresistance<br />

No extraordinary magnetoresistance was observed in the entire<br />

temperature range. The magnetoresistance at 200 K and 300 K is a few percent<br />

in a 70 kOe field and approximately proportional to H 2 (Figure 5-7). Reliable<br />

resistivity data could be measured on the crystal sample down to 40 K. No<br />

discontinuity in the resistivity was found near T C .<br />

5. 2. 2 Small Polaron Hopping<br />

The electrical resistivity of Gd 0.67 Ca 0.33 MnO 3 is consistent with adiabatic<br />

small polaron hopping conductivity (section 4.2.6)R h Texp(E a /k B T) [58, 68]<br />

R h = 2k B /3ne 2 a 2 ν particularly at high temperatures. Here, k B is Boltzmann’s<br />

constant, e is the electronic charge, n is the number density of charge carriers<br />

— about 0.7 carrier per Mn, a is the site-to-site hopping distance, and ν is the<br />

longitudinal optical phonon frequency. This behavior has also been observed<br />

in the La containing compound described in section 4.2.2 [103, 146].<br />

Resistivity data are often displayed on an Arrhenius plot (ln(ρ) vs. 1/T) which<br />

will give a straight line for a band gap semiconductor [15, 147]. The slightly<br />

different temperature dependence of the data provides the significant<br />

nonlinearity in such a fit (Figure 5-6). A polaron hopping activation energy,<br />

E a , of 1940K, and an attempt frequency of 8 × 10 13 s -1 is inferred. The small<br />

polaron mechanism implies that unlike a bandgap semiconductor, the highly<br />

temperature dependent quantity is the mobility, not the carrier concentration.<br />

113


114 Chapter 5<br />

A high, temperature independent carrier concentration is consistent with<br />

Hall effect measurements (sect 4.2.4) on La 0.67 Ca 0.33 MnO 3 .<br />

5. 2. 3 Variable Range Hopping<br />

The resistivity of the manganites has also been attributed to variable range<br />

hopping due to Anderson localization [148]. At temperatures less than 150K,<br />

this picture produces a somewhat better fit to our data (Figure 5-7), but is<br />

clearly inferior to the polaron model at high temperatures (Figure 5-6). The<br />

fit to R 0 exp(T 0 /T) 1/4 gives T 0 = 1.7 × 10 9 K. In the theory of variable range<br />

hopping [71, 149], k B T 0 ≈ 21/(ζ 3 N(E F )) where ζ is the decay length of the<br />

localized wave function, and N(E F ) is the density of localized states at the<br />

Fermi level. For ζ ≈ a = 3.9Å, the distance between neighboring Mn atoms,<br />

this implies N(E F ) ≈ 2.4 × 10 18 eV -1 cm -3 which is typical for disorder<br />

semiconductors but about 1500 times less than the density of states found in<br />

the specific heat of manganites which become metallic at low temperatures<br />

(≈ 5 mJ mol -1 K -2 )[124].<br />

5.3 X-ray Absorption Fine Structure<br />

The primary quantity in XAFS data analyses is χ(k)=(µ(k)-µ 0 (k))/µ 0 (k)),<br />

where k is the ejected photoelectron wave vector, µ(k) is the total absorption<br />

due to the absorbing atomic species, and µ 0 (k)) is the portion of µ(k) that does<br />

not include the photoelectron backscattering off neighboring atoms. This<br />

backscattering causes an interference at the absorbing atom which is manifest<br />

as oscillations in χ(k). A Fourier transform of kχ(k) thus produces peaks that<br />

correspond to the distribution of atoms around the absorbing atom.<br />

Figure 5-8 shows these Fourier transforms.


Local Structure and Magnetism in Gd 0.67 Ca 0.33 MnO 3<br />

Figure 5-8 Fourier transform of kχ(k) from (a) Mn K-edge and<br />

(b) Gd L III -edge data on Gd 0.67 Ca 0.33 MnO 3 . The solid lines are<br />

data collected at T = 69 K, while the triangles (∆) are data<br />

collected at T = 40 K. Agreement between data above and<br />

below T C is well within the errors of the experiment.<br />

Transform ranges for the Gd edge data are from 3.5-12.5 Å -1<br />

and Gaussian broadened by 0.3 Å -1 . Transform ranges for the<br />

Mn edge data are from 3.2-12.5 Å -1 and Gaussian broadened by<br />

0.3 Å -1 .<br />

The agreement between the data above and below T C is very good and<br />

places limits on any atom-position changes around either the Gd or M n<br />

atoms within ±0.005 Å, and changes in the Debye-Waller broadening factor to<br />

±0.002 Å (Pnma; a = 5.52Å, b = 7.50Å, c = 5.34Å).<br />

115


116 Chapter 5<br />

5. 3. 1 Relationship of structure to CMR<br />

Unlike the colossal magnetoresistance manganites, Gd 0.67 Ca 0.33 MnO 3 does<br />

not have a metal-insulator transition or a large magnetoresistance near the<br />

Curie temperature. The negative magnetoresistance is only of order 1% in<br />

large fields and is proportional to H 2 (or M 2 since it scales with χ 2 ) like that<br />

observed in La 0.67 Ca 0.33 MnO 3 in the paramagnetic state (section 7.2.1) [150].<br />

Thus, Gd 0.67 Ca 0.33 MnO 3 is clearly in the class of 1/3 doped manganites which,<br />

due to the small size of Gd 3+ [100], change from paramagnetic insulators to<br />

ferromagnetic insulators at T C . There is much interest in finding a structural<br />

change at T C in the giant magnetoresistive manganites, since it is predicted to<br />

have a large effect on the resistive transition [11]. There is now evidence [151-<br />

154] for a decrease in structural disorder as the GMR manganites become<br />

ferromagnetic metals. Gd 0.67 Ca 0.33 MnO 3 shows the contrapositive: there is no<br />

discontinuity in the atomic disorder for a material which does not become<br />

metallic at T C . This has also been noted in manganites with different dopant<br />

concentrations: La 0.88 Ca 0.12 MnO 3 [151] and La 0.5 Ca 0.5 MnO 3 [154].<br />

5.4 Conclusion<br />

Gd 0.67 Ca 0.33 MnO 3 is ferrimagnetic with a compensation temperature of<br />

about 15K due to the interaction and ordering of the Gd 3+ ions. The<br />

qualitative features of the magnetic properties can be accounted for with a<br />

simple two sublattice (Gd and Mn) molecular field model. The large high<br />

field susceptibility at low temperatures may indicate a canting of the magnetic<br />

sublattices. Contrary to the conclusion in [136] that there does not exist long-<br />

range magnetic order (only spin-glass magnetism) in the ferromagnetic-<br />

insulator region of the tolerance-factor phase-diagram, Gd0.67Ca0.33MnO3 shows definite long-range order and a compensation point which can not be<br />

explained by spin-glass phenomena alone. The resistivity is consistent with<br />

small polaron conductivity over a broad temperature range, with a possible<br />

crossover to a different mechanism such as variable-range hopping at low


Local Structure and Magnetism in Gd 0.67 Ca 0.33 MnO 3<br />

temperatures. There is no noticeable change in the structure, as determined<br />

by XAFS, or conductivity at the ferrimagnetic transition. Recent theoretical<br />

and experimental work conclude that structural effects at T C are associated<br />

with the metal-insulator transition and large magnetoresistance. Since<br />

Gd 0.67 Ca 0.33 MnO 3 is in the doping regime where a metal-insulator transition<br />

and large magnetoresistance are not observed at T C , this work supports that<br />

conclusion, i. e. that a significant structural change is associated not with the<br />

ferromagnetic transition, but the metal-insulator transition.<br />

117


6. Magnetoconductivity in La 0.67Ca 0.33MnO 3<br />

In this chapter, the nature of the field dependence of the resistivity in<br />

La 0.67 Ca 0.33 MnO 3 - or magnetoresistance above and below T C is studied in more<br />

detail. For this study, sample LCM 17 (Chapter 4, Table 4-2) was used since it<br />

shows magnetic and transport properties consistent with bulk samples.<br />

Measurements at 0.9T C = 237 K and 1.1T C = 289 K will have considerable<br />

magnetoresistance while being clearly either in the ferromagnetic or<br />

paramagnetic state. In order to achieve maximum temperature stability, data<br />

were collected only after there was no appreciable drift in the resistivity. The<br />

major conclusions have been reproduced with other samples which like this<br />

one have high T C ’s and no domain boundary magnetoresistance (section<br />

4.2.8). The work presented in this chapter has been published in [150].<br />

Two configurations were used: 1) magnetic field parallel to the film,<br />

which was used for longitudinal and transverse magnetoresistance analysis<br />

since there is almost no demagnetization field. Since the van der Pauw<br />

configuration is used, the longitudinal magnetoresistance has a small<br />

component due to transverse magnetoresistance and vice versa. 2) magnetic<br />

field perpendicular to the film, used primarily for Hall effect measurements.<br />

6.1 Anisotropic magnetoresistance<br />

The magnetoresistance at 289 K (1.1T C ) on increasing and decreasing<br />

magnetic field is displayed in Figure 6-1 for longitudinal and transverse<br />

directions. The two configurations give identical magnetoresistance.<br />

The high field (|H| > 1 kOe) magnetoresistance at 0.9T C is also shown in<br />

Figure 6-1 for both increasing and decreasing fields. There is no appreciable<br />

hysteresis for |H| > 200 Oe. The transverse and longitudinal<br />

magnetoresistance in high fields are nearly identical when the same current<br />

path is used. Different current paths however, give a slightly different<br />

118


Magnetoconductivity in La 0.67 Ca 0.33 MnO 3<br />

magnetoresistance at 0.9T C indicating some magnetic/electronic<br />

inhomogeneity below T C . All current paths have the functional form of the<br />

magnetoresistance described below, and therefore the apparent<br />

inhomogeneity does not affect the conclusions of this paper. The transverse<br />

magnetoresistance with the field perpendicular to the film is identical to the<br />

transverse magnetoresistance with the field parallel to the film with an<br />

additional demagnetization field of (1750+/-100 Oe); this was used to estimate<br />

M at 0.9 T C .<br />

6.2 Magnetoresistance models<br />

In order to assess the potential of the colossal magnetoresistance in a<br />

magnetic field sensor, (such as the read head in a magnetic recording device)<br />

knowledge of the magnitude and exact field dependence of the<br />

magnetoresistance is critical, primarily in low magnetic fields, H.<br />

Experimental data generally show that for low fields the magnetoresistance is<br />

approximately quadratic in H above the Curie temperature, T C and is more<br />

cusp like or linear below T C [116]. There exists, however, many counter<br />

examples to this generalization [22, 25, 114] presumably due to inhomogenous<br />

samples where T C is not so well defined [116]. The best empirical model of the<br />

resistivity (ρ = ρ 0 e M/M 0) around T C [114] gives a linear field dependence. Since<br />

the earliest reports [4], the magnetoresistance has been claimed or assumed to<br />

be independent of the angle between the magnetic field and the current [113].<br />

The microscopic mechanism of the magnetoresistance can be quite<br />

complicated and still not accurately predict much of the observed<br />

magnetoresistance phenomena. Much of the theory assumes an electron<br />

band transport mechanism [55], which may not be appropriate in the case of<br />

the manganites because they are not metallic above T C . Some microscopic<br />

models of transport in the manganites have predicted the conductivity σ or<br />

119


120 Chapter 6<br />

resistivity ρ ∝ M 2 [113, 155, 156] (which is isotropic) with limited experimental<br />

verification.<br />

6. 2. 1 General Model<br />

The difference above and below T C as well as the relationship to the<br />

magnetization M can be shown more generally by only considering the<br />

symmetry of the resistivity tensor - with no assumptions about or even a<br />

reference to microscopic transport mechanisms. Following Landau and<br />

Lifshitz [157], the relation between the electric field E and the current density J<br />

is given by the resistivity tensor ρ: E i = ρ ik J k where the components of ρ are<br />

functions of H and M. Since H and M are vectors which are antisymmetric<br />

under time reversal, ρ must have the following symmetry: ρ ik (H,M) =<br />

ρ ki (-H,M) = ρ ki (H,-M) = ρ ik (-H,-M). The symmetric, s ik , and antisymmetric, a ik ,<br />

parts of ρ ik = s ik + a ik then have the following properties: s ik (H,M) = s ik (-H,M) =<br />

s ik (H,-M) and a ik (H,M) = -a ik (-H,M) = -a ik (H,-M). Thus the components of s ik are<br />

even functions of H and M while those of a ik are odd functions.<br />

The measured resistivity ρ, the diagonal component of the resistivity<br />

tensor, contains no contribution from a ik and therefore must be even in H and<br />

M. Assuming the resistivity is an analytic function with respect to H and M<br />

(which themselves may not be analytic in T), the first terms in the expansion<br />

of ρ in powers of H and M is then ρ = ρ 0 + αM 2 + βH 2 + γH•M. At this point,<br />

one can already expect that for T > T C (M=0) the low field magnetoresistance<br />

should be proportional to H 2 while for T < T C (M≠0) the term linear in H may<br />

be expected to dominate. Clearly a negative magnetoresistance can not obey<br />

this relation for arbitrarily large H, for the resistivity would eventually<br />

become negative; higher order terms or inverse powers of H are then<br />

required. The above argument is as valid for the conductivity tensor σ, where<br />

J i = σ ik E k , as it is for ρ. In the case of the manganites, the magnetoconductivity


Magnetoconductivity in La 0.67 Ca 0.33 MnO 3<br />

is positive and unlike the magnetoresistance may not require further<br />

expansion in H.<br />

6.2.1.1 Magnetoconductivity model<br />

For an isotropic conductor the longitudinal and transverse<br />

magnetoconductances (and magnetoresistances) are, in general, not equal.<br />

The general form of the relation between J and E in an isotropic conductor<br />

[157] up to terms quadratic in H and M is:<br />

J = σ 0 E + α 1 M 2 E + α 2 (E•M)M +<br />

β 1 H 2 E + β 2 (E•H)H +<br />

γ 1 (H•M)E + γ 2 (M•E)H + γ 3 (E•H)M + (1)<br />

σ H H×E + σ A M×E<br />

If all the coefficients (except α 1 ) are allowed to be functions of M 2 , (e.g.<br />

σ 0 → σ 0 + α 1 M 2 + … ) then (1) is a complete expansion in H up to H 2 and all<br />

powers of M.<br />

The β i terms provide the M = 0 magnetoconductance quadratic in H.<br />

When β 2 = 0, the longitudinal and transverse magnetoconductances are<br />

equal. Similarly, the α 2 term leads to magnetoconductance depending on the<br />

angle between E and M - known as anisotropic magnetoresistance 6.1. The γ i<br />

provide magnetoconductance linear in H only when M ≠ 0. The equation<br />

used to fit the data is formally equivalent to a circuit containing a resistor (ρ ∞ )<br />

in series with the magnetoconductor (e.g. σ(H) = σ 0 + σ Η 2H 2 ), although the<br />

physical significance of such an equivalent circuit is not entirely clear. In any<br />

event, ρ ∞ is presumably the intrinsic spin independent resistivity found to be<br />

proportional to a constant plus a T 2 term, described in sections 4.2.1 and 7.2.2<br />

[103].<br />

121


122 Chapter 6<br />

The Hall effects are contained in the antisymmetric part of σ and ρ and<br />

therefore can be expanded in odd powers of H and M. The first two terms in<br />

the Hall resistivity is R H H×J + R A M×J, where R H is the normal Hall coefficient<br />

and R A is the anomalous Hall coefficient. Since there is no evidence to prefer<br />

the Hall conductivity of equation 1 (≈ -R H /H 2 ) to the Hall resistivity, the Hall<br />

effect is analyzed using R H and R A .<br />

It is possible that the observed magnetoconductivity could be reduced to<br />

terms involving only M 2 and (E•M)M since M ≈ M 0 + χH gives the same field<br />

dependence described above. This would imply a strict relationship among<br />

the various β i (T) and γ i (T) which could be determined by simultaneous<br />

measurements of M and the magnetoconductivity. The resistivity saturates<br />

in a magnetic field [113, 158] more rapidly than does the magnetization<br />

(giving M 2 behavior only at small M), indicating again that it is probably the<br />

conductivity not the resistivity which is proportional to M 2 . This is discussed<br />

in more detail in chapter 7.<br />

The reciprocal nature of the conductivity and resistivity leads to only an<br />

abstract distinction between them. Thus, the simpler, empirical description of<br />

the resistivity in terms of a magnetoconductivity may be accidental.<br />

Nevertheless, it warrants further consideration. The standard additive<br />

scattering time interpretation of the magnetoresistivity requires a more<br />

complex H dependence to describe the suppression of magnetic scattering by a<br />

magnetic field. The simplicity of the magnetoconductivity expression may<br />

signify the basis of an alternative transport interpretation. Taken literally, the<br />

magnetic field appears to open parallel channels of conduction with a simple,<br />

physical functional dependence.


MagnetoResistance ∆R/R(0) (%)<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

6.2.1.1.1 T > T C regime<br />

Magnetoconductivity in La 0.67 Ca 0.33 MnO 3<br />

ρ ∞<br />

1/βH 2 or 1/γ|Η|<br />

1/σ 0<br />

La 0.67 Ca 0.33 MnO 3<br />

T = 0.9Tc<br />

T = 1.1Tc<br />

-60<br />

-80 -60 -40 -20 0 20 40 60 80<br />

Magnetic Field (kOe)<br />

Figure 6-1 High field (longitudinal) magnetoresistance above<br />

and below T C for La 0.67 Ca 0.33 MnO 3 film. The solid lines show<br />

the fit using the indicated equivalent circuit<br />

The 1.1T C data fit very well with three parameters (ρ ∞ = 1.3 mΩcm, σ ο =<br />

0.14 (mΩcm) -1 and σ Η 2 = 43 × 10 -6 (mΩcm) -1 (kOe) -2 ) to ρ(H) = ρ ∞ + 1/(σ 0 + βH 2 ).<br />

A much less satisfactory fit is obtained with ρ(H) = ρ(0) + aH 2 + bH 4 .<br />

Above T C , M = 0 (or at least M = χH which gives the same field<br />

dependence of σ) so the diagonal element of the isotropic magneto-<br />

conductance tensor σ(H) = σ 0 + β 1 H 2 + β 2 (E•H) 2 /E 2 . From the fit described<br />

above at 1.1 T C, β 1 = 43 × 10 -6 (mΩcm) -1 (kOe) -2 and β 2 ≈ 0 is estimated.<br />

123


124 Chapter 6<br />

∆R/R (%)<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

Transverse<br />

Sum<br />

Longitudinal<br />

-600 -400 -200 0 200 400 600 -1<br />

Field (Oersted)<br />

Magnetization<br />

Figure 6-2 Low field magnetoresistance and magnetization<br />

(relative units) of La 0.67 Ca 0.33 MnO 3 at 0.9 T C . The sum of the<br />

longitudinal and transverse resistances minimizes the effect<br />

of the anisotropic magnetoresistance.<br />

6.2.1.1.2 T < T C regime<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Relative<br />

Magnetization<br />

The longitudinal data fit well with three parameters (ρ ∞ = 0.58 mΩcm, σ ο =<br />

0.35 (mΩcm) -1 and σ Η = 0.013 (mΩcm kOe) -1 ) to ρ(H) = ρ ∞ + 1/(σ ο + σ Η |H|). A<br />

significantly less satisfactory fit is obtained with an exponential form [114,<br />

156]: ρ(H) = ρ ∞ + ae -|H|/b .<br />

Below T C and for M ≠ 0 parallel to H, i. e. M = MH/|H|, then up to terms<br />

linear in H, σ(H) reduces to (σ 0 + α 1 M 2 ) + α 2 M 2 (E•H) 2 /E 2 H 2 + γ 1 M|H| +<br />

(γ 2 + γ 3 )M|H|(E•H) 2 /E 2 H 2 . By fitting the 0.9T C data, γ 1 = σ Η /M = 73 × 10 -6 (mΩcm


Magnetoconductivity in La 0.67 Ca 0.33 MnO 3<br />

kOe emu) -1 is estimated. For equal longitudinal and transverse<br />

magnetoresistance, γ 2 + γ 3 ≈ 0.<br />

The |H| form of the magnetoresistance below T C provides a cusp in the<br />

magnetoresistance near H = 0. This is in contrast to the rounded curve<br />

observed above T C .<br />

6.2.1.1.3 Anisotropic magnetoresistance<br />

The low field data are shown in Figure 6-2 with the average of the<br />

transverse and longitudinal magnetoresistances. The vertical offset of these<br />

data is approximate, i.e. it is calculated from measurements of a different,<br />

patterned sample. The switching of the resistivity at +/-65 Oe is due to the<br />

switching of the magnetization (coercivity). This can be seen in Figure 6-2<br />

where the magnetization, relative to the saturated 0.9 T C value of about<br />

1µ B /Mn, is shown for comparison.<br />

The anisotropic magnetoresistance (AMR) at 0.9T C appears to provide the<br />

nonlinear magnetoresistance in low fields, (Figure 6-2) aside from the<br />

magnetization switching at 65 Oe. Since 1/(ρ trans + ρ long ) ≈ α 2 (E trans •M) 2 +<br />

α 2 (E long •M) 2 ≈ α 2 M 2 (Cos 2 (θ) + Cos 2 (θ + 90°)) = α 2 M 2 , the sum of the transverse<br />

and longitudinal resistances should be independent of the direction of M,<br />

leaving only the magnetoresistance linear in H (Figure 6-2). This is<br />

approximately what happens. The larger (smaller) than linear peak<br />

(depression) in the longitudinal (transverse) magnetoresistance for H < 100<br />

Oe is apparently due to AMR since the sum produces a curve which looks<br />

like |H| that switches at the coercive field. Similar AMR has been<br />

independently verified elsewhere [159].<br />

125


126 Chapter 6<br />

R(H)-R(-H) (µΩ cm)<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

La 0.67 Ca 0.33 MnO 3 Hall Effect<br />

T = 1.1 T C<br />

T = 0.9 T C<br />

0.0<br />

0 10 20 30 40 50 60 70<br />

Magnetic Field (kOe)<br />

Figure 6-3 Hall effect of La 0.67 Ca 0.33 MnO 3 below (fully<br />

magnetized data only) and above T C .<br />

6.2.1.1.4 Hall Effect<br />

The Hall effect (Figure 6-3) is calculated from R(H) - R(-H), which should<br />

remove all contributions to the transverse voltage due to magnetoresistance<br />

and changes in magnetic homogeneity. The Hall effect at 0.9T C shows a clear<br />

contribution at H=0 due to the anomalous Hall effect [160] as well as a term<br />

linear in H. From this linear term, a carrier density of n = 0.85 holes/cell is<br />

calculated which is the same as that observed at 5 K (section 4.2.4), where no<br />

anomalous Hall effect was detected. The anomalous Hall effect apparently<br />

increases with temperature, which is quite unexpected since it is usually<br />

proportional to M [160]. Nevertheless, this has also been observed in the<br />

magnetic spinels such as CuCr 2 Se 4 [132-134]. At 1.1 T C the Hall effect has an


Magnetoconductivity in La 0.67 Ca 0.33 MnO 3<br />

electron-like slope at low fields that slowly reverses to hole-like in high fields.<br />

The low field slope can be attributed to the anomalous Hall effect resulting<br />

from the magnetization due to the Curie-Weiss susceptibility, χ: R hall = R A M +<br />

R H H = (χR A + R H )H. Thus while the sample magnetizes the Hall effect looks<br />

electron-like, but once the magnetization saturates, the Hall effect returns to<br />

hole-like.<br />

6.3 Conclusion<br />

In summary, an accurate and simple model, based solely on symmetry<br />

without reference to any mechanism, is proposed for the magnetoresistance<br />

in La 0.67 Ca 0.33 MnO 3 . One can also conclude from this model that the<br />

mechanism of the magnetotransport is best described by a<br />

magnetoconductance, and therefore mechanistic theories should predict at<br />

least a M 2 dependence of the conductivity. The functional form and estimate<br />

of parameters presented above should not only help guide the development<br />

of a mechanistic theory, but also provide a basis for predicting the<br />

magnetotransport properties when designing a device.<br />

In homogeneous materials, it is clear that the magnetoresistance<br />

maximizes near the ferromagnetic Curie temperature T C and rapidly<br />

decreases at lower or higher temperatures. The magnetoresistance also slowly<br />

saturates as the magnetic field is increased past several Tesla. Thus, the<br />

conductivity behaves much like the magnetization: above T C the<br />

conductivity and magnetization are low, while below T C the conductivity and<br />

magnetization rapidly increase. Furthermore, near T C both the conductivity<br />

and resistivity can be increased by an applied magnetic field. The theory of<br />

double exchange, which has been developed to explain these properties of the<br />

manganites [7, 112, 155], predicts this correlation between the conductivity and<br />

the magnetization.<br />

127


7. Critical Transport and Magnetization of La 0.67Ca 0.33MnO 3<br />

In homogeneous La 0.67 Ca 0.33 MnO 3 , it is clear that the magnetoresistance<br />

maximizes near the ferromagnetic Curie temperature T C and rapidly<br />

decreases at lower or higher temperatures. The magnetoresistance also slowly<br />

saturates as the magnetic field is increased past several Tesla. Thus, the<br />

conductivity behaves much like the magnetization: above T C the<br />

conductivity and magnetization are low, while below T C the conductivity and<br />

magnetization rapidly increase. Furthermore, near T C both the conductivity<br />

and magnetic moment can be increased by an applied magnetic field. The<br />

theory of double exchange, which has been developed to explain these<br />

properties of the manganites [7, 112, 155], predicts this correlation between the<br />

conductivity and the magnetization.<br />

The relationship between the conductivity (or resistivity) and the<br />

magnetization has been examined both experimentally and theoretically.<br />

Several theoretical models [31, 112, 121, 155, 156, 161-164] predict an M 2<br />

dependence of the resistivity (or conductivity). The first term of the Taylor<br />

series expansion of the resistivity (or conductivity) in terms of M should be<br />

M 2 for all models due to symmetry considerations [150].<br />

Experimentally, a correlation between the resistivity and magnetization<br />

can be found by plotting one as a function of the other [24, 114, 115, 121, 158,<br />

165]. These plots show that the resistivity is roughly proportional to the<br />

square of the magnetization M 2 for small M. For larger M however, the<br />

resistance decreases more slowly in relationship to M 2 as if it were saturating<br />

before the magnetization does. Thus the relationship between the resistivity<br />

and the magnetization is more complicated than ρ ∝ M 2 . This relationship<br />

may instead be well described by an M 2 dependence of the conductivity in a<br />

slightly more complicated circuit [150]. In the previous chapter, only the<br />

128


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

magnetic field H dependence of the resistivity is shown to be more fully<br />

described in terms of a magnetoconductive circuit, ρ = ρ ∞ + 1/(σ 0 + σ H 2H 2 ) for<br />

T > T C and ρ = ρ ∞ + 1/(σ 0 + σ H |H|) for T < T C , where ρ ∞ , σ 0 , σ H 2 and σ H are<br />

parameters. In order to determine if the magnetoconductivity, σ H 2H 2 and<br />

σ H |H| can replaced by a common term proportional to M 2 as suggested above,<br />

the temperature dependencies and magnitudes of σ H 2, σ H and M need to be<br />

examined.<br />

Alternatively, it has been found in several cases that the correlation<br />

ρ ∝ exp[-M(H,T)/M E ] is a good fit to the data [24, 114, 115]. This formulation<br />

does not yet have a good theoretical understanding.<br />

In this chapter, which will be published separately, the temperature and<br />

field dependence of the magnetization and magnetoresistance, above and<br />

below T C are reported. A well annealed La 0.67 Ca 0.33 MnO 3 thin film on a<br />

LaAlO 3 substrate grown by MOCVD was used for transport measurements as<br />

described in chapter 4. DC conductivity measurements were performed using<br />

the method of Van der Pauw [52, 53]. Only the longitudinal<br />

magnetoresistance data are shown. The transverse magnetoresistance is<br />

nearly identical to the longitudinal as was discussed in chapter 6. After<br />

stabilizing the temperature, repeated resistance R vs. H curves were measured<br />

and fit with three parameters (ρ ∞ , σ 0 , and σ H 2 or σ H ) to ρ = ρ ∞ + 1/(σ 0 + σ H 2H 2 )<br />

for T > T C and ρ = ρ ∞ + 1/(σ 0 + σ H |H|) for T < T C . The data fit well to these<br />

forms except the few degrees near T C (where a combination of the two forms<br />

is better) and at low temperatures. For T < 50 K the magnetoresistance shows<br />

2<br />

no sign of saturating (ρ = [ρ∞ + 1/σ0 ] - σH |H|/σ0 ), so only two parameters<br />

2<br />

([ρ∞ + 1/σ0 ] and σH /σ0 ) can be extracted.<br />

129


130 Chapter 7<br />

2 )<br />

M 2 (µ B<br />

The focus of this study is on the isotropic properties. The effects due to the<br />

anisotropic magnetoresistance (section 6.1) and magnetocrystalline anisotropy<br />

are minimized by concentrating on the H > 1000 Oe data for T < T C . Even in<br />

this preliminary investigation, peculiar magnetic properties are found that<br />

require a more complex physical model than that used for a typical<br />

ferromagnet.<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

La 0.67 Ca 0.33 MnO 3 Crystal Arrott Plot<br />

254K<br />

258K<br />

260K<br />

261K<br />

262K = Tc<br />

263K<br />

264K<br />

265K<br />

266K<br />

268K<br />

0.0<br />

0 5000 10000 15000 20000<br />

H/M (Oe/µ )<br />

B<br />

7.1 Magnetism near T C<br />

Magnetization measurements were performed while the magnetic field H<br />

was decreasing. Data for increasing H field are identical to the H decreasing<br />

data for T > T C . Polycrystalline pellets were prepared (section 2.1.1) and this<br />

material was used to for floating-zone laser-heated crystal growth described i n<br />

M (µ B /Mn)<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 5 10<br />

Magnetic Field (kOe)<br />

15<br />

270K<br />

La 0.67 Ca 0.33 MnO 3 Pellet 274K<br />

M = χH + χ 3 H 3<br />

χ 3 > 0<br />

272K<br />

M = χH<br />

274K 276K<br />

Figure 7-1. M 2 vs. H/M plot for La 0.67 Ca 0.33 MnO 3 float zone<br />

crystal. A mean field ferromagnet has linear isotherms with a<br />

positive slope. The negative slope for T >T C indicates a faster<br />

than linear increase in M (inset) due to a highly unusual<br />

positive non-linear susceptibility χ 3 .


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

2 )<br />

M 2 × |T/T C - 1| -2β (µ B<br />

section 2.1.2. The float zone product (referred to below as “crystal”) is dense<br />

and crystallographically highly-oriented. The polycrystalline pellets and a<br />

float zone crystal used for magnetization measurements have very similar<br />

magnetic properties. A demagnetization factor N (H d = 4πNM) of about 0.5<br />

was estimated and used to correct the data. The demagnetization correction is<br />

large only for H small. The conclusions of this work are unaffected by<br />

variations in N.<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

La 0.67 Ca 0.33 MnO 3 Scaled Arrott Plot<br />

β = 0.27<br />

γ = 0.90<br />

T C =262.2K<br />

T < T C<br />

T > T C<br />

0<br />

0 100 200 300 400 500 600 700 800<br />

H/M × |T/T - 1|<br />

C -γ (kOe/µ )<br />

B<br />

Figure 7-2. Data from Figure 7-1 (using the same symbols)<br />

scaled with β = 0.27 and γ = 0.90. According to the scaling<br />

hypothesis, all the T < T C data should lie on a single curve<br />

while the T > T C data should lie on a separate, single curve.<br />

Most of the M(H, T) data can be scaled [90] with β = 0.27 and γ = 0.90.<br />

According to the scaling hypothesis, the magnetic equation of state in the<br />

critical region depends only on the scaled variables H/|T C /T - 1| β+γ and<br />

M/|T C /T - 1| β . A plot of the scaled M 2 and scaled H/M, shown in Figure 7-2,<br />

will then have only two curves: one branch for the T < T C data and another<br />

131


132 Chapter 7<br />

for T > T C . The regions of negative slope on the M 2 vs. H/M plot (χ 3 > 0)<br />

make the scaled T > T C curve also approach M 2 = 0 with a negative slope. For<br />

T - T C < 0.01 T C , the scaled data break away from this region of negative slope<br />

to giving a nearly straight curve typical of a ferromagnet. This crossover in<br />

scaling suggests that the true critical regime only begins with |T - T C | < 0.01<br />

T C , and that the exponents reported here, which will be quite useful in<br />

modeling the magnetization, are effective rather than true critical exponents.<br />

The true critical regime is reached when the magnetic critical fluctuations<br />

become as large as the net magnetization itself.<br />

The experimental value of the critical exponent δ, H = M δ when T = T C ,<br />

depends strongly on the demagnetization correction, and therefore not<br />

analyzed here. The scaling relation δ = 1 + γ/β can be used to estimate δ. The<br />

other critical exponents γ and β do not vary significantly when a different<br />

demagnetization correction is used.<br />

7. 1. 1 Spontaneous magnetization exponent<br />

The square of the magnetization M 2 , plotted vs. H/M (Figure 7-1)<br />

facilitates understanding the critical behavior of La 0.67 Ca 0.33 MnO 3 . The<br />

isotherm which extrapolates to M 2 = 0, H/M = 0 is the critical isotherm T = T C .<br />

In this way T C = 262.2 K ± 0.5 K is estimated. The uncertainty in T C is due to<br />

the uncertainty of the demagnetization correction. The isotherms below T C<br />

should be approximately linear and intersect H/M = 0 at M 0 . From these<br />

M 0 (T) the (magnetic order parameter) critical exponent β ≈ 0.30 and<br />

T C = 262.2 K can be estimated by fitting M 0 (T) ∝ (1 - T/T C ) β (Figure 7-3).<br />

The fit is not as good as this method, in the same apparatus, allows<br />

(Appendix A), perhaps due to the contributions in low fields from the<br />

magnetocrystalline anisotropy, which have been ignored. Therefore, the


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

values for β reported here are more uncertain than those reported for other<br />

materials.<br />

M 0 (µ B /Mn)<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

ln(M)<br />

1.0<br />

0.5<br />

0.0<br />

-0.5<br />

M ∝ (Tc-T) β<br />

β = 0.30<br />

-1.0<br />

-2 -1 0 1 2 3 4 5<br />

ln(262.2-T)<br />

0.0<br />

190 200 210 220 230 240 250 260 270<br />

Temperature (K)<br />

Figure 7-3. Saturation Magnetization, M 0 as a function of<br />

temperature for La 0.67 Ca 0.33 MnO 3 crystal. At each temperature,<br />

the value shown is M extrapolated to H = 0 as given by the<br />

intercept in Figure 7-1.Solid line is fit to M 0 (T) ∝ (1 - T/T C ) β<br />

with β = 0.30.<br />

Typical experimental values for the critical exponent β in Fe, Ni and YIG<br />

[92] are 0.37 ± 0.02, which are near the theoretical values (Ising β = 0.33,<br />

Heisenberg β = 0.36). The related metallic ferromagnets SrRuO 3 (Appendix A)<br />

and La 0.5 Sr 0.5 CoO 3 [92] have β ≈ 0.36 and β = 0.361 respectively. Previous<br />

magnetic data [166, 167] on La 0.67 Ca 0.33 MnO 3 have been fit with such values of<br />

β, suggesting that the critical region has been reached. For example, muon-<br />

133


134 Chapter 7<br />

spin-relaxation measurements [167] give β = 0.345 ± 0.015 with T C = 274 K.<br />

The double layer perovskite La 1.4 Sr 1.6 Mn 2 O 7 has a small [168] β = 0.25 possibly<br />

due to 2-D ferromagnetism. The significance of the low values for β given<br />

here for La 0.67 Ca 0.33 MnO 3 is unknown due to the uncertainty of the<br />

measurement.<br />

7. 1. 2 Susceptibility exponent<br />

The T > TC isotherms should intersect M 2 = 0 at 1/χ(T, H = 0) = 1/χ0 ∝<br />

(T/T C - 1) γ and can be used to give an effective γ ≈ 0.7 and T C = 263 K ± 1 K.<br />

This value of γ is unexpected particularly since it is less than one (Figure 7-4).<br />

For a typical ferromagnet, the plot of 1/χ vs. T is close to linear at high<br />

temperatures and then become slightly concave up as T approaches T C . In<br />

this way, the high temperature, linear extrapolation of 1/χ to 1/χ = 0 gives the<br />

paramagnetic Curie temperature, Θ P ≈ 257 K. The ferromagnetic Curie<br />

temperature, T C is the temperature where 1/χ and M 0 vanish, and normally<br />

T C < Θ P . However, for γ < 1, the plot of 1/χ vs. T is concave down and Θ P < T C .<br />

When critical fluctuations are included in the theory, γ increases due to<br />

the suppression of the transition (Ising γ = 1.24, Heisenberg γ = 1.39). Typical<br />

experimental values for γ (Fe, Ni and YIG [92] γ = 1.2 ± 0.2; SrRuO 3 (Appendix<br />

A) γ = 1.17) are greater than one.<br />

An inhomogeneous ferromagnet, with a range of T C ‘s, can provide such a<br />

negative curvature resulting in an apparent γ < 1; however, this will also<br />

make M 0 decrease more linearly (β ≈ 1) as T approaches T C from below. Since<br />

the observed β is not larger but smaller than expected, a range of T C ‘s is<br />

unlikely. In ferrimagnets such a negative curvature is expected due to the


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

competition between ferromagnetic and antiferromagnetic interactions. As<br />

pointed out by DeGennes, there is competition between double exchange<br />

(ferromagnetic) and superexchange (antiferromagnetic) as will be discussed<br />

further in section 7.1.4. One of the early theories of double exchange<br />

presented by Anderson and Hasagawa [7] also predicts a downward curvature<br />

in 1/χ vs. T.<br />

It is not too surprising that two different values of γ (0.7 and 0.9) are<br />

obtained from the same data. The two methods weight the data differently<br />

and as discussed above, the true critical region where γ is a constant has<br />

probably not been reached.<br />

7. 1. 3 Positive nonlinear susceptibility<br />

A peculiarity of the T > T C isotherms of Figure 7-1, which likely affects the<br />

determination of γ, is their negative slope as they approach M 2 = 0. This slope<br />

is proportional to the nonlinear susceptibility. For M = χH + χ 3 H 3 , χ is the<br />

linear susceptibility and χ 3 is the third-order nonlinear magnetic<br />

susceptibility. The slope of a T > T C isotherm in an M 2 vs. H/M Arrott plot as<br />

M 2 approaches zero is given by -χ 4 /χ 3 . For a normal ferromagnet, χ > 0 while<br />

χ 3 < 0 giving a positive slope for all T > T C isotherms. However in<br />

La 0.67 Ca 0.33 MnO 3 a negative slope is found which implies that χ 3 > 0. This can<br />

also be seen in a plot of M vs. H, which is shown in Figure 7-1. A positive χ 3<br />

produces the faster than linear increase in M with H which leads to an<br />

inflection point in the M vs. H curve due to the saturation of M for large H.<br />

This effect is not likely due to the sample preparation since it is seen in both<br />

the polycrystal and float-zone crystal samples. Arrott plots of SrRuO 3<br />

(Appendix A) at similar temperatures do not show χ 3 > 0, which indicates that<br />

the effect is not due to the measurement system. A more pronounced<br />

135


136 Chapter 7<br />

1/χ (10kOe/µ B )<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

260 270 280 290 300<br />

inflection point has been observed in the double layer perovskite manganite<br />

La 1.4 Sr 1.6 Mn 2 O 7 above T C [169].<br />

1/χ ∝ (T-T C ) γ<br />

γ = 0.7<br />

Temperature (K)<br />

A positive χ 3 can be observed in a ferromagnet or spin glass during<br />

irreversible magnetization processes; for example, while increasing the<br />

magnetic field from a demagnetized state. However, once the material is<br />

magnetized at a high field, the curve of M vs. H for decreasing field almost<br />

always has a negative χ 3 . The data in the inset of Figure 7-1 includes data for<br />

both increasing and decreasing fields, showing no signs of hysteresis [170].<br />

ln(1/χ)<br />

11<br />

10<br />

9<br />

1 2 3<br />

ln(263K-T)<br />

Figure 7-4. Inverse magnetic susceptibility, 1/χ 0 as a function<br />

of temperature for La 0.67 Ca 0.33 MnO 3 crystal. At each<br />

temperature, the value shown is H/M extrapolated to H = 0 as<br />

given by the intercept in Figure 7-1. Solid line is fit to 1/χ 0 ∝<br />

(T/T C - 1) γ with γ = 0.7 and T C = 263K.


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

Thus a positive χ 3 due to the irreversibility of a spin glass seems highly<br />

unlikely. More plausible explanations will be discussed in section 7.1.4.<br />

7. 1. 4 Additional magnetic interaction<br />

The unusual γ < 1 may be related to χ3 > 0. Both effects suggest an<br />

additional mechanism. There could, for instance, be additional magnetic<br />

interactions, such as an antiferromagnetic component to the coupling. A<br />

canted-antiferromagnet ground-state is expected theoretically from the<br />

double-exchange mechanism in conjunction with super-exchange [8], and has<br />

been found by neutron diffraction [139, 143] in the related compound<br />

Pr 0.67 Ca 0.33 MnO 3 . If there is a competition between antiferromagnetic and<br />

ferromagnetic fluctuations in the paramagnetic state, the application of a<br />

magnetic field will favor ferromagnetism and thus induce a larger magnetic<br />

moment. This would explain the observed positive χ 3 . The corresponding<br />

transition in the ordered state is called a metamagnetic transition. Competing<br />

magnetic interactions can also cause the downward curvature in 1/χ vs. T<br />

seen in Figure 7-4. This is commonly observed in a ferrimagnet (e.g.<br />

Gd 0.67 Ca 0.33 MnO 3 described in section 5.1.2) where some strong interactions are<br />

antiferromagnetic, but near T C a ferromagnetic susceptibility is observed. The<br />

canting of the moments by as much as 45° would still give 92% of the<br />

expected net moment, which is about that observed (section 4.1) at low<br />

temperatures for La 0.67 Ca 0.33 MnO 3 .<br />

A different but more general explanation for the unusual magnetic<br />

behavior observed in La 0.67 Ca 0.33 MnO 3 requires some additional mechanism<br />

which provides the stronger than mean field behavior, i.e. the ferromagnetic<br />

interaction strength appears to increase, beyond that predicted by the mean<br />

field approximation, as the magnetization increases. For instance, as T<br />

approaches T C , the strength of the interaction increases, effectively increasing<br />

T C . This causes the susceptibility to diverge faster than expected, providing<br />

137


138 Chapter 7<br />

γ < 1. Also, at a constant T > T C , as H is applied M increases, which induces<br />

stronger coupling (higher effective T C ) and therefore increasing the<br />

susceptibility in a nonlinear way: χ 3 > 0.<br />

This additional mechanism may be related to the ferromagnetic metal to<br />

paramagnetic insulator phase transition. La 0.67 Ca 0.33 MnO 3 , like other<br />

materials which exhibit a metal-insulator phase transition, undergoes a static<br />

and even dynamic structural transition [151, 154, 171, 172]. The changes in the<br />

electronic and atomic structure should alter the magnetic coupling, i.e. the<br />

exchange J is not a constant, but a function of M or ρ. Specifically, since the<br />

metallic state has a lower volume than the insulating state, one might expect<br />

the magnetic coupling to strengthen as the material becomes more metallic.<br />

The metallic character, conversely, is clearly associated with the<br />

magnetism as is predicted in the double exchange model. As the applied<br />

magnetic field (and hence the magnetization) increases, the material becomes<br />

more metallic. Not only does the resistivity decrease, but the metal-insulator<br />

transition (defined by a change in sign of dρ/dT) is pushed to higher<br />

temperatures. If the induced metallic state promotes a further increase in the<br />

magnetization, a positive nonlinear susceptibility (χ 3 > 0) and a rapidly<br />

diverging susceptibility (γ < 1) may be generated. Thus, the structural change<br />

may cause the magnetism to increase faster than that expected from the mean<br />

field model.<br />

La 0.67 Ca 0.33 MnO 3 may have a smaller critical region than a normal<br />

ferromagnet because of these additional interactions. In typical ferromagnets,<br />

short range exchange interactions result in a short coherence length of the<br />

magnetic fluctuations. These fluctuations have low energy and can be<br />

observed relatively far from T C . In La 0.67 Ca 0.33 MnO 3 additional electronic and<br />

lattice interactions may increase the fluctuation coherence length. With


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

Magnetization (µ B /Mn)<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

more volume involved for each fluctuation, they will have higher energy<br />

and only observable very near T C .<br />

La 0.33 Ca 0.67 MnO 3 Magnetization<br />

0.9 T C<br />

1.1 T C<br />

0.0<br />

0 10 20 30 40 50 60<br />

Internal Field (kOe)<br />

Figure 7-5. Magnetization in a magnetic field for a<br />

La 0.67 Ca 0.33 MnO 3 polycrystalline pellet at 0.9 and 1.1 T C . The<br />

solid lines indicate the linear regions in each case.<br />

7.2 Magnetoresistance<br />

To illustrate the need for the magnetoconductivity expression described i n<br />

section 6.2.1.1 for the magnetoresistance, the two temperatures 0.9 T C and<br />

1.1T C are considered first. The magnetization of a polycrystalline pellet at<br />

these two temperatures is shown in Figure 7-5. At T = 0.9 T C , M ≈ M 0 + χH<br />

except in fields less than a few kOe when the material is not yet magnetically<br />

saturated. At T = 1.1 T C , M ≈ χH is a good approximation particularly for<br />

H < 40 kOe. The films used in this study have been shown to have similar<br />

magnetic behavior as the bulk samples (chapter 4), but due to their small<br />

volume and the complication of the substrate provide less accurate magnetic<br />

139


140 Chapter 7<br />

∆R/R (%)<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

-60<br />

ρ ∞<br />

∆R/R<br />

1/σ H |H|<br />

1/σ 0<br />

0.9 T C La 0.33 Ca 0.67 MnO 3<br />

-M 2<br />

-60 -40 -20 0 20 40 60<br />

Internal Field (kOe)<br />

data. Therefore, in what follows, the magnetotransport on films is compared<br />

with magnetization measured on bulk (ceramic and crystal) samples.<br />

The magnetoresistance, ∆R/R, and M 2 as functions of internal field<br />

(making the demagnetization correction described above) are compared below<br />

and above T C in Figure 7-6 and Figure 7-7. They are clearly related<br />

qualitatively, although there are marked differences at high fields as<br />

saturation is approached, particularly for T < T C .<br />

The saturation of the magnetoresistance is best described by the fit to a<br />

magnetoconductivity circuit (section 6.2.1.1), as if the magnetic field were<br />

opening channels of conductivity. This circuit assumes a linear M vs. H<br />

∆R/R (%)<br />

-50<br />

-55<br />

σ = σ 0 + σ H |H|<br />

σ = A e |H|/B<br />

-60<br />

40 50 60 70<br />

Magnetic Field (kOe)<br />

-4.5<br />

-5.0<br />

-5.5<br />

-6.0<br />

-6.5<br />

-7.0<br />

Figure 7-6. Magnetoresistance of La 0.67 Ca 0.33 MnO 3 film<br />

compared with -M 2 of a pellet, both at 0.9 T C . The solid line for<br />

the magnetoresistance data shows the fit using the indicated<br />

equivalent circuit. The dashed line in the inset compares the<br />

exponential fit.<br />

-M 2 (µ B<br />

2 /Mn)


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

(section 6.2.1.1). The resistivity is fit to ρ = ρ ∞ + 1/σ where σ = σ 0 + σ H 2H 2 for<br />

T > T C , while for T < T C , σ = σ 0 + σ H |H| is used. Each fit has three free<br />

parameters ρ ∞ , σ 0 , and σ H 2 or σ H . If the magnetoconductivity is indeed<br />

proportional to M 2 as suggested above, then the nonlinearity of M vs. H<br />

(Figure 7-5) will slightly alter the fitting parameters, particularly ρ ∞ .<br />

∆R/R (%)<br />

0<br />

-8<br />

-16<br />

-24<br />

-32<br />

-40<br />

-48<br />

1.1 T C La 0.33 Ca 0.67 MnO 3<br />

∆R/R -M 2<br />

ρ ∞<br />

1/σ H H 2<br />

2<br />

1/σ 0<br />

-60 -40 -20 0 20 40 60<br />

Internal Field (kOe)<br />

Figure 7-7. Magnetoresistance of La 0.67 Ca 0.33 MnO 3 film<br />

compared with -M 2 of a pellet, both at 1.1 T C . The solid line for<br />

the magnetoresistance data shows the fit using the indicated<br />

equivalent circuit.<br />

The exponential dependence of the resistivity first proposed by Hundley et<br />

al. [114] on the magnetization ρ = ρ E exp[-M(H,T)/M E ] does not fit the<br />

magnetoresistance data as well above T C as the fit described above. For T > T C<br />

and H small, ρ ≈ ρ E exp[-χH/M E ] ≈ ρ E (1 - χH/M E ) predicts a magnetoresistance<br />

linear in H while the observed magnetoresistance is quadratic (Figure 7-7).<br />

Below T C the exponential fit is not at good as the magnetoconductivity circuit<br />

0.0<br />

-0.5<br />

-1.0<br />

-1.5<br />

-2.0<br />

-2.5<br />

-M 2 (µ B<br />

2 /Mn)<br />

141


142 Chapter 7<br />

(Figure 7-6), in so far as the H dependence is concerned; however, the<br />

exponential dependence upon M is found, as will be shown below.<br />

ρ ∞ (mΩcm)<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

ρ ∞<br />

σ 0 and ρ ∞ above T C<br />

280 290 300 310 320 330<br />

Temperature (K)<br />

Figure 7-8. Fitting parameters σ 0 and ρ ∞ for T > T C in a<br />

La 0.67 Ca 0.33 MnO 3 film. The temperature dependence of these<br />

two parameters reflect the insulating behavior of the material.<br />

7. 2. 1 Magnetoresistance scaling above T C<br />

The magnetoresistance data at various temperatures above T C were fit to<br />

the expression ρ = ρ ∞ + 1/(σ 0 + σ H 2H 2 ). The resulting parameters σ 0 and ρ ∞ are<br />

shown in Figure 7-8. The parameter σ 0 remains relatively constant, slightly<br />

increasing as expected for a semiconductor. The significance of ρ ∞ will be<br />

discussed below.<br />

σ 0<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

σ 0 (mΩcm -1 )


σ (10<br />

H -10 mΩcm -1 Oe -2 2<br />

)<br />

Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

ln(σ 2)<br />

Η<br />

-22<br />

-23<br />

-24<br />

-25<br />

-26<br />

280 300 320 340 360<br />

Temperature (K)<br />

The parameter σ H 2, which determines the magnetoresistance, is highly<br />

temperature dependent (Figure 7-9), diverging at T C . The data can be fit with<br />

a critical exponent of -2.0 ± 0.1: σ H 2(T) = σ H 2(2T C ) (T/T C - 1) -2 , where T C = 262 K ±<br />

1 K and σ H 2(2T C ) ≈ 8 × 10 -13 Oe -2 mΩcm -1 . This result supports an M 2<br />

dependence of the magnetoconductance (or magnetoresistance) since<br />

M 2 ∝ (T/T C - 1) -2γ above T C , where, as experimentally determined above,<br />

γ ≈ 0.9. Specifically, it is found that σ(H,T) = σ 0 + σ H 2(T)H 2 ≈ σ 0 + σ M 2M 2 . Using<br />

the Curie Weiss susceptibility from [103], M = χH ≈ 5 × 10 -4 (emu G -1 cm -3 ) ×<br />

(T/T C - 1) -1 H, gives σ M 2 = 3 × 10 -6 Oe -2 mΩcm -1 .<br />

Slope = -2<br />

σ ∝ χ<br />

Η 2<br />

2<br />

Above T<br />

C<br />

Resistance ≈ 1/(1+M 2 )<br />

2.8 3.2 3.6 4.0 4.4<br />

ln(T-262K)<br />

Figure 7-9. Fitting parameter σ H 2 as a function of temperature<br />

in a La 0.67 Ca 0.33 MnO 3 film for T > T C . The temperature<br />

dependence of σ H 2 and the square of the susceptibility are the<br />

same, indicating a relationship between the<br />

magnetoconductance and M 2 .<br />

143


144 Chapter 7<br />

The related form ρ ∝ -M 2 would also be consistent with our low field data<br />

since in low fields ∆ρ = -H 2 2<br />

σH 2/σ0 and σ0 is nearly constant with respect to<br />

temperature. This form was also observed above T C in La 1-x Sr x CoO 3 [165] and<br />

in small fields for lower T C films [158] of La 0.7 Ca 0.3 MnO 3 . In the temperature<br />

range of this experiment, T > 1.01 T C , 4πM < H, so B = H + 4πM ≈ H and<br />

therefore not very temperature dependent. Thus the resistivity or<br />

conductivity is not proportional to B or B 2 .<br />

7. 2. 2 Magnetoresistance scaling below T C<br />

Below T C , ρ ∞ can be interpreted as due to scattering processes which would<br />

persist if the moments were perfectly ordered. As shown in Figure 7-10, the<br />

resistivity in a infinite field, ρ ∞ , follows the A + BT 2 fit to the low-<br />

temperature, H = 0, intrinsic resistivity described in section 4.2.1. In a true<br />

infinite field (H = ∞) one would expect the resistance to be continuous with a<br />

positive (metallic) slope as T increases through T C . While there is a factor of 2<br />

discrepancy between ρ ∞ (T C ) extrapolated from above (Figure 7-8) and below<br />

(Figure 7-10) this is not unreasonable because two different fitting equations<br />

are used. Particularly above T C , the fitting parameter ρ ∞ may not reflect the<br />

true H = ∞ resistivity because the nonlinearity of M upon H is not taken into<br />

account.<br />

The parameter σ 0 , which can be related to the resistivity in zero field<br />

ρ(H = 0) ≈ ρ ∞ + 1/σ 0 , describes the limiting field-dependent, spin-disorder<br />

scattering processes present as H approaches 0. Unlike ρ ∞ , 1/σ 0 diverges as the<br />

Curie temperature is approached. A critical exponent of about 1.8 (Figure 7-<br />

10) is found: σ 0 (T) ≈ σ 0 (0) (1 - T/T C ) 1.8 , where σ 0 (0) ≈ 2 × 10 -3 mΩcm -1 .


ρ ∞ and 1/σ 0 (mΩ cm)<br />

Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

ln(σ 0 )<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

σ 0 ∝ (T C -T) 1.8<br />

-2<br />

2 3 4 5<br />

ln(262K-T)<br />

σ 0 and ρ ∞ below T C<br />

0.0<br />

80 100 120 140 160 180 200 220 240<br />

Temperature (K)<br />

Figure 7-10. Fitting parameters σ 0 and ρ ∞ for T < T C in a<br />

La 0.67 Ca 0.33 MnO 3 film. ρ ∞ is governed by the A + BT 2 terms in<br />

the resistivity while σ 0 diverges at T C . The inset shows σ 0 data<br />

fit with a (T C - T) 1.8 power law (dashed line), and σ ∝<br />

exp(M/M E ) (solid line). The zero field resistivity ρ(H = 0) =<br />

ρ ∞ + 1/σ 0 is shown for comparison.<br />

The parameter σ H is more difficult to interpret than the other parameters.<br />

2<br />

Although the magnetoresistance, ∆R/HR proportional to σH /σ0 , increases as<br />

T C is approached, σ H decreases. Thus the large magnetoresistance found<br />

below T C is due to the divergence of 1/σ 0 and not σ H . The parameter σ H does<br />

not always appear to vanish completely as T approaches T C , making it difficult<br />

to fit to a (1 - T/T C ) n power law. Assuming T C = 262K (determined from the<br />

critical properties of σ H 2 and σ 0 ) the exponent for σ H is about 0.7.<br />

ρ ∞<br />

1/σ 0<br />

ρ(H = 0)<br />

A + BT 2 fit<br />

145


146 Chapter 7<br />

σ H (10 -5 mΩcm -1 Oe -1 )<br />

20<br />

15<br />

10<br />

5<br />

0<br />

ln( σ H )<br />

100 150 200 250<br />

Temperature (K)<br />

Figure 7-11. Fitting parameter σ H as a function of temperature<br />

in a La 0.67 Ca 0.33 MnO 3 film for T < T C . The solid line shows the<br />

best fit to the data using a critical exponent of 0.7.<br />

There is no simple relationship between either the magnetoconductance<br />

or the magnetoresistance with M 2 below T C because the measured scaling<br />

exponents do not agree with those predicted from an M 2 dependence as<br />

discussed in the following. Shown in Figure 7-5 is the magnetization below<br />

T C which can be approximated with M(H) = M 0 + χH , and therefore<br />

M 2 2<br />

≈ M 0 + 2χM0H. Assuming σ(H,T) = σ0 + σHH ≈ σM 2 M(H,T) 2 as suggested<br />

2<br />

above, then below TC the following relations should hold: σ0 = σM 2M0 and<br />

σ H = 2χσ M 2M 0 . The temperature dependence of M 0 and χ below T C should be<br />

governed by the scaling exponents β and γ, via M 0 ∝ (1 - T/T C ) β and<br />

χ ∝ (1 - T/T C ) -γ where experimentally β = 0.3 and γ ≈ 0.9 (typical critical<br />

exponents are β ≈ 0.35 and γ ≈ 1.2). Thus one would expect the critical<br />

-9<br />

-10<br />

-11<br />

3.0 4.0 5.0<br />

ln(262K-T)


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

exponent for σ 0 and σ H to be 0.6 and -0.6 respectively. Experimentally,<br />

however σ 0 ∝ (1 - T/T C ) 1.8 and σ H ∝ (1 - T/T C ) 0.7 .<br />

An M 2 dependence of the magnetoresistance (as opposed to the<br />

magnetoconductance) where ρ ≈ ρ0 - ρMM 2 2<br />

≈ (ρ0 - ρMM0 ) - 2ρMχM 0H below TC ,<br />

2<br />

is also inconsistent. For low fields ρ = (ρ∞ + 1/σ0 ) - (σH /σ0 )|H| is found<br />

2<br />

experimentally, which has a field dependent term (σH /σ0 ), which varies as<br />

(1 - T/T C ) -2.9 . This is quite different from 2ρ M χM 0 which has a critical exponent<br />

of β - γ ≈ -0.6.<br />

Below T C , in agreement with previous work [23, 107], the conductivity is<br />

exponentially dependent upon M. Removing the slowly varying<br />

contribution due to ρ ∞ , this model predicts σ(H,T) = σ E exp[M(H,T)/M E ] ≈<br />

σ 0 + σ H H. For H = 0 this reduces to σ 0 = σ E exp[M 0 (T)/M E ]. The experimental<br />

relationship M 0 (T) ∝ (1 - T/T C ) β with β = 0.3 can be used to fit the magneto-<br />

resistance data. The higher quality of this fit compared to the M 2 fit is shown<br />

in Figure 7-10, with M E ≈ 0.4 µ B (M E = 1.0 µ B in [107]) and σ E ≈ 4 × 10 -3 mΩcm -1 .<br />

An exponential dependence of the conductivity may suggest a tunneling<br />

mechanism is responsible. Tunneling conductivity depends exponentially on<br />

the length of the tunneling barrier. If, in some way, this barrier is decreased<br />

by an increase in the magnetization, then the conductivity will depend<br />

exponentially on M as observed for large M. Spin dependent tunneling is<br />

reported to be the mechanism of the large domain boundary<br />

magnetoresistance observed in these materials [140].<br />

Furthermore, the temperature dependence of σ H can also be explained<br />

with the exponential model. According to this model, the field dependent<br />

conductivity<br />

d d M M<br />

σH σ σE<br />

dH dH<br />

σσ E σE σχ<br />

e<br />

/ E<br />

= = 0 dM<br />

= = 0<br />

M dH M<br />

(where<br />

E<br />

E<br />

147


148 Chapter 7<br />

σ E exp[M 0 (T)/M E ] ≈ σ 0 is used) should have the same temperature dependence<br />

as σ 0 χ. Experimentally, σ 0 χ has an approximate temperature dependence of<br />

(1 - T/T C ) 1.8-0.9 , assuming γ ′ (susceptibility exponent for T < T C ) = γ. This is<br />

within experimental error of the critical exponent found (Figure 7-11) for σ H<br />

(about 0.7).<br />

7. 2. 3 Magnetoresistance scaling at T C<br />

At or very near T C , the magnetic properties should be described by the<br />

critical exponent δ, where M ∝ H 1/δ when T = T C . The magnetoresistance data<br />

for T = 262 K ≈ T C is shown in Figure 7-12, and can be fit to ρ = ρ ∞ +<br />

1/(σ 0 + σ H nH n ), where n is an additional free parameter. The data fit well with<br />

n = 1.2 ± 0.1, where the variation of n arises from different ranges and<br />

weighting of the fit. Thus, if at T C , the conductivity σ ∝ M 2 then σ ∝ H 2/δ , or<br />

δ = 2/n. This is clearly not the case since 2/n ≈ 1.6 while typically δ ≈ 4.8, or<br />

from the data presented above and the scaling relation δ = 1 + γ/β the value<br />

δ ≈ 4.3 is found.<br />

The T = 262 K data fit well to the composite relation proposed by Sun et al.<br />

[23] σ ∝ M 2 exp[M/M E ] combined with H = M δ . Above T C this relation reduces<br />

to σ ∝ M 2 , which, as was shown above, is quite accurate. Below T C , the<br />

exponential term dominates invalidating the simple σ ∝ M 2 relationship.


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

∆R/R (%)<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

-60<br />

-70<br />

-80<br />

ρ ∞<br />

262 K ≈ T C La 0.33 Ca 0.67 MnO 3<br />

n = 1.3<br />

1/σ H H n<br />

n<br />

1/σ 0<br />

-60 -40 -20 0<br />

Field (kOe)<br />

20 40 60<br />

Figure 7-12. Magnetoresistance of La 0.67 Ca 0.33 MnO 3 film at 262<br />

K ≈ T C . The solid line shows the fit (for the full data on a<br />

linear scale) using the indicated equivalent circuit.<br />

7. 2. 4 Relation to low temperature magnetoresistance<br />

This qualitative correlation between the magnetization and the resistivity<br />

may also account for the intrinsic, small, linear magnetoresistance (Figure 4-<br />

8) found even at the lowest temperatures where the magnetization is nearly<br />

saturated. If at these low temperatures σ = σ 0 + σ M 2M 2 then ∆σ/σ (or ∆ρ/ρ) is<br />

given by ∆σ = 2χσ M 2M 0 H. Since ∆σ = -∆ρ/ρ 2 , the susceptibility χ required to<br />

give the observed magnetoresistance is χ = -∆ρ/(2ρ 2 σ M 2M 0 H ) = 2.5 × 10 -4<br />

emu/Oe cm 3 using the results from chapter 4 for ∆ρ/H = -1.5 × 10 -8 mΩcm/Oe,<br />

ρ = 0.125 mΩcm, and M 0 = 3.4 µ B /Mn = 550 emu/cm 3 . This is within a factor<br />

of three of the observed value (section 4.2.9, Figure 4-8) χ = 9 × 10 -5<br />

emu/Oe cm 3 . Similarly acceptable, is the exponential model<br />

149


150 Chapter 7<br />

σ = σ E exp[M(H,T)/M E ], which gives at low T and H,<br />

∆ρ/H = -χexp[-M 0 /M E ]/M E σ E ; requiring χ = 32.9 × 10 -5 emu/Oe cm 3 . A mag-<br />

2 2 2<br />

netoresistive model ρ = (ρ∞ + 1/σ0 ) - (σM 2/σ0 )M giving ∆ρ = -2χσM2M0H /σ0 is<br />

2<br />

significantly worse since it would predict χ = -∆ρσ0 /(2σM2M0H ) = 2 × 10 -7<br />

emu/Oe cm 3 .<br />

In a similar manner, the correlation between the magnetization and the<br />

resistivity should contribute to the T 2 contribution (section 4.2.1.2) to the low<br />

temperature resistivity ρ = R 0 + R 2 T 2 + … . This is because the magnetization<br />

can be approximated with M = M 0 (1 - (T/Θ) 2 ) for these temperatures (where<br />

Θ ≈ 500 K, section 4.1.1), giving ρ(T) ∝ -M(T) ∝ T 2 . For the σ = σ 0 + σ M 2M 2<br />

model ρ(T) - ρ(0) = -ρ 2 (σ(T) - σ(0)) = 2σM 2ρ 2 2 2 -7 -2 2<br />

M0 (T/Θ) ≈ 1.26 × 10 mΩcm K T .<br />

For the ρ = exp[-M(H,T)/M E ]/σ E model ρ(T) - ρ(0) = exp[-M 0 /M E ]/σ E ×<br />

(T/Θ) 2 M 0 /M E ≈ 1.14 × 10 -6 mΩcm K -2 T 2 . Both of these proposed contributions<br />

are less than the observed value (section 4.2.1.2) R 2 ≈ 1.96 × 10 -5 mΩcm K -2 ,<br />

suggesting that scattering mechanisms such as electron-electron, or electron-<br />

magnon scattering dominates the R 2 T 2 term.<br />

7.3 Conclusion<br />

A scaling relation has been found that satisfactorily characterizes the<br />

magnetic and magnetoresistive properties of La 0.67 Ca 0.33 MnO 3 near T C .<br />

However the scaling exponents, particularly γ < 1, show that critical scaling<br />

has not been reached even though the measurements reach 0.01 T/T C where<br />

simple ferromagnets are expected to exhibit critical scaling.<br />

An unusual positive nonlinear magnetic susceptibility, which does not<br />

disrupt the scaling, is also observed. The positive non-linear susceptibility<br />

and susceptibility exponent γ < 1 may result from the coupling of the


Critical Transport and Magnetization of La 0.67 Ca 0.33 MnO 3<br />

ferromagnetic and metal-insulator transitions, making the exchange J a<br />

function of M.<br />

The scaling of both the H and T dependence of the magnetoconductance<br />

above T C is consistent with a simple M 2 dependence of the conductivity.<br />

However, below or very near T C this correlation predicts the observed H<br />

dependence but can not account for the T dependence. To fit these data, the<br />

model for the magnetoresistance needs to have an exponential dependence<br />

on the magnetization for large M, which reduces to M 2 for small M , such as<br />

that previously suggested by Sun et al. [23] σ ∝ M 2 exp[M/M E ].<br />

151


Appendix A. Critical Behavior and Anisotropy in Single<br />

Crystal SrRuO 3<br />

Introduction<br />

Strontium ruthenate, SrRuO 3 , has many physical properties which make<br />

it unique among perovskite oxides. First, it is metallic in the undoped state<br />

[36, 126, 173-176]. This is even more unusual since the ruthenium in SrRuO 3<br />

is in a high oxidation state; whereas many other metallic perovskites must be<br />

formed in reducing environments and therefore unstable in air at high<br />

temperatures. The remarkable chemical stability and simple chemical<br />

formula makes metallic SrRuO 3 quite attractive for use in epitaxial thin film<br />

heterostructures with other perovskite oxides when metallic layers are<br />

desired. Indeed, the nearly cubic SrRuO 3 [177] is often preferred over the<br />

more distorted CaRuO 3 when making structures such as electrodes for<br />

ferroelectrics or superconductor-normal metal-superconductor junctions.<br />

The other striking feature of SrRuO 3 is its ferromagnetism with a<br />

reasonably high transition temperature (163 K) and large saturation moment<br />

(> 1µ B ) [174, 175, 178, 179]. SrRuO 3 is the only ferromagnetic perovskite oxide<br />

of a 4d or 5d transition metal. Moreover, SrRuO 3 has the largest saturation<br />

moment known to arise from 4d electrons, making it more related to the iron<br />

group ferromagnetic metals (Fe, Co and Ni) than to the weak itinerant<br />

electron ferromagnets such as ZrZn 2 . SrRuO 3 also has very strong cubic<br />

magnetic anisotropy, requiring magnetic fields in excess of 10 Tesla to saturate<br />

the magnetization in the hard directions. Such a strong anisotropy makes<br />

measuring even the simplest properties, such as saturation magnetization,<br />

difficult. It is the purpose of the present work to determine the magnetic<br />

properties of SrRuO 3 by measurements of magnetically-soft single crystals<br />

along the easy magnetic direction.<br />

152


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

There have been several studies of SrRuO 3 in the past three decades,<br />

mostly on polycrystalline samples which are quite easy to prepare. SrRuO 3<br />

has a very slightly distorted (GdFeO 3 type) perovskite structure. The<br />

deviation from perfect cubic perovskite is so small that it has often been<br />

undetected. The resistivity of polycrystalline, and epitaxial thin film SrRuO 3<br />

shows a cusp in dρ/dT at the ferromagnetic Curie temperature T C . This is<br />

commonly observed for metallic ferromagnets and is attributed to spin<br />

disorder scattering [125]. Reported T C 's tend to vary from 150K to 165K.<br />

The saturation magnetization of SrRuO 3 has been both difficult to<br />

measure and interpret. Low spin Ru 4+ in an octahedral coordination has four<br />

4d electrons in the t 2g triply degenerate state, giving two paired and two<br />

unpaired electrons. Since the orbital component of angular momentum J<br />

will be quenched, J = S = 1 is expected. The measured Curie constant of the<br />

paramagnetic state is consistent with this model (expected: µ eff = √2gJ(J + 1) µ B<br />

= 2.83 µ B ; measured = 2.67 µ B [175]). For a localized moment ferromagnet, the<br />

saturation magnetization M S is predicted to be M S = gJ µ B = 2.0 µ B for SrRuO 3 .<br />

Measured values of M S are much less. Polycrystalline SrRuO 3 reaches about<br />

0.85 µ B [175, 178, 179] in low fields but continues to increase in higher<br />

magnetic fields. At 125 kOe it was noted that M had reached 1.55 µ B but had<br />

not yet saturated [178]. Early neutron diffraction derived a moment of 1.4 ±<br />

0.4 µ B [178] with no evidence for any antiferromagnetic order (spin canting).<br />

Recent theoretical investigations predict incomplete band splitting and a large<br />

moment of about 1.6 µ B [176, 180]. Early explanations for the low value of M S<br />

included spin canting, band magnetism, and incomplete alignment of the<br />

magnetization due to magnetocrystalline anisotropy [178]. The present work<br />

shows that SrRuO 3 has a large saturation moment of 1.6 µ B as predicted by<br />

these calculations.<br />

Single crystals of SrRuO 3 can be grown from a SrCl 2 flux [36]. The<br />

resistivity of such crystals is consistent with the results on polycrystalline<br />

153


154 Appendix A<br />

samples. Magnetization and magnetic torque measurements of single crystals<br />

[181-183] showed that SrRuO 3 has a high cubic anisotropy with 〈110〉 (cubic<br />

cell) being the easy direction, with nearly square hysteresis loops. Some<br />

previous measurements reported for single crystals are questionable, for<br />

example, the measured value of M S = 1.1 µ B . In this work new magnetization<br />

data is shown to clarify these points.<br />

Experimental<br />

Single crystals of SrRuO 3 were grown by slow cooling in a SrCl 2 flux [36,<br />

184]. Polycrystalline SrRuO 3 was prepared from stoichiometric quantities of<br />

SrCO 3 and Ru metal repeatedly reacted at 1260°C, and was used as the source<br />

material for crystal growth. The SrCl 2 was dried in air at 110°C. A mixture<br />

with approximate weight ratio 1:20 of SrRuO 3 :SrCl 2 was melted in a platinum<br />

crucible with lid at 1260°C for 94 hrs. The sample was cooled to 800°C at 1°/hr<br />

and then to room temperature at ~40°/hr. Crystals of SrRuO 3 less than 1 m m<br />

in diameter were found at the bottom of the crucible after removing the flux<br />

with water. Most crystals were cubo-octahedron shaped and grew with a<br />

3-fold symmetric axis (presumably [111]) perpendicular to the Pt surface. The<br />

(cubic) crystal orientation was determined by the symmetry of the faces. The<br />

actual orthorhombic symmetry [177] was confirmed by powder x-ray<br />

diffraction and Transmission Electron Microscopy.<br />

Magnetization in fields up to 70 kOe was measured using a Quantum<br />

Design MPMSR 2 SQUID magnetometer. The samples were attached to a<br />

plastic straw with a small amount of Apiezon N vacuum grease. Samples<br />

were oriented on the straw visually using crystal faces which had obvious 2-,<br />

3- and 4-fold rotational symmetry. At room temperature, the crystal can<br />

rotate in a large field to align a paramagnetic easy direction with the field.<br />

This was used as a final adjustment when aligning along the easy 〈110〉


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

directions. At low temperatures, the grease solidifies so the sample cannot<br />

rotate. The accuracy of the magnetization measurement was estimated by<br />

measuring a sphere of yttrium iron garnet described in section 3.2.1.1. For the<br />

measure-ments reported here, the sample holder was fixed at the rotation<br />

angle which gave the maximum magnetization. Several crystals were<br />

measured.<br />

Results<br />

Magnetic Moment (µ B /Ru)<br />

1.7<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

M S<br />

√(2/3) M S<br />

SrRuO 3 Single Crystal<br />

[110]<br />

[1-10]<br />

[101]<br />

[111]<br />

[100]<br />

1.2<br />

1.1<br />

0<br />

M /√2<br />

S<br />

10 20 30<br />

-2<br />

-75<br />

40<br />

-50 -25 0 25 50<br />

H (kOe)<br />

50 60<br />

75<br />

70<br />

Magnetic Field (kOersted)<br />

Figure A- 1. Magnetization at 5 K of SrRuO 3 single crystal<br />

along several crystallographic directions showing strong cubic<br />

but not uniaxial magnetocrystalline anisotropy. Inset shows<br />

the full hysteresis loop of the single crystal data along with<br />

that of a polycrystalline pellet for comparison.<br />

Figure A- 1 shows hysteresis loops at 5 K for SrRuO 3 single crystal i n<br />

various crystallographic directions and a polycrystalline pellet. The crystals<br />

have a low coercive field (≈ 10 Oe) compared to that of the polycrystalline<br />

pellet (3000 Oe). But more importantly, the crystals show very little hysteresis<br />

M (µ B /Ru)<br />

2<br />

1<br />

0<br />

-1<br />

Pellet<br />

155


156 Appendix A<br />

while the pellet displays noticeable hysteresis even in fields greater than<br />

40 kOe.<br />

The rapid, linear approach to saturation (with respect to the applied field)<br />

found in all directions can be attributed to demagnetization. Until the sample<br />

becomes fully magnetized, the demagnetization field H d is equal to the<br />

applied field H a resulting in an internal field of zero (H i = H a - H d ). From this<br />

slope (M = H a /4πN), one can calculate the demagnetization factor N and<br />

therefore calculate the internal field. The measured demagnetization factors<br />

N, (H d = 4πNM) are 0.25, 0.28, 0.31, 0.49, 0.66 for the [110], [1-10], [101], [100] and<br />

[111] directions respectively. These seem reasonable considering the shape of<br />

the crystals. Since a uniaxial magnetocrystalline anisotropy will also give a<br />

linear increase in M if H is applied along the hard direction, it is difficult to<br />

distinguish it from the effect of the demagnetization field in this study. Since<br />

the demagnetization field can be as large as a thousand Oersted, one can only<br />

conclude that the uniaxial anisotropy field is less than a thousand Oersted,<br />

which is considerably less than that ( > 50 kOe) reported previously [182].<br />

In the [110] direction SrRuO 3 rapidly approaches saturation and then<br />

remains relatively constant (square hysteresis loop), as is expected for a<br />

magnet with a magnetic field along the easy direction. The crystal was also<br />

measured in the [1-10] and [101] directions which would be equivalent by<br />

symmetry to the [110] if the crystals were cubic. Since the magnetization is the<br />

same along these 〈110〉 type directions, it can be concluded that the magnetic<br />

properties of single crystal SrRuO 3 are essentially cubic, i.e. only cubic<br />

magnetocrystalline anisotropy is detected. Beyond this initial saturation<br />

along the easy 〈110〉 directions, there is a small but measurable increase in the<br />

magnetization which is linear in magnetic field and has a slope of 6 × 10 -7<br />

µ B /Oe.


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

2 )<br />

M 2 (µ B<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

SrRuO 3 Crystal Arrott Plot<br />

140K<br />

145K<br />

152K<br />

Log(H)<br />

156K<br />

158K<br />

2<br />

160K 1<br />

162K<br />

-0.9<br />

163K = Tc<br />

-0.8 -0.7 -0.6 -0.5<br />

Log(M)<br />

-0.4 -0.3<br />

164K<br />

166K<br />

168K<br />

170K<br />

0.0<br />

0 5000 10000 15000<br />

H/M (Oe/µ )<br />

B<br />

20000 25000 30000<br />

The H = 0, T = 0 saturation moment M S found along the easy 〈110〉<br />

directions are about 1.62 µ B /Ru. In the remnant state (H reduced to zero), the<br />

magnetization should lie along the nearest easy direction. Thus the expected<br />

remnant (H = 0) magnetization along the [100] or [111] direction is simply the<br />

cosine of the angle it makes with the closest 〈110〉 direction. For 〈100〉 the<br />

expected remnant magnetization is M S /√2, and for 〈111〉 directions √(2/3)M S is<br />

expected. These are extremely close to the experimental values (Figure A- 1).<br />

In the other primary directions, there is a nonlinear approach to satur-<br />

ation which is characteristic of materials with cubic anisotropy. The magneto-<br />

crystalline anisotropy constants K 1 and K 2 can be estimated from these<br />

magnetization curves [185]. The magnetocrystalline anisotropy energy E can<br />

be defined in terms of the cosine of the angle M makes with the three crystal<br />

2 2 2 2 2 2<br />

axes: αi = xi • M/M. For a cubic material E = K1 (α1 α2 + α2 α3 + α3 α1 ) +<br />

4<br />

3<br />

T = 163K H ∝ M δ<br />

δ = 4.2(1)<br />

Figure A- 2. Arrott Plot of SrRuO 3 single crystal along easy<br />

[110] direction. Inset, critical isotherm (T = 163K ≈ T C ) on a log<br />

scale fit to M δ ∝ H with δ = 4.2.<br />

157


158 Appendix A<br />

M 0 (µ B )<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Log(M 0 )<br />

2 2 2<br />

K2 (α1 α2 α3 ) + … . In SrRuO3 the magnetic easy axes are 〈110〉, which requires<br />

K 1 < 0. Similarly, 〈100〉 are the hard axes (〈111〉 are intermediate) which<br />

implies 2.25|K 1 | < K 2 < 9|K 1 |. The [100] magnetization should intersect that<br />

of the easy axis [110] at H a = -2K 1 /M S . An extrapolation of the [100] M vs. H<br />

curve gives -2K 1 /M S ≈ 109 kOe. Alternatively, the area between the [100] and<br />

the [110] M vs. H curves should be equal to -K 1 /4. This method gives a value<br />

for -2K 1 /M S ≈ 96 kOe. The area method can also be used with the [111] curve<br />

to estimate 2K 2 /M S ≈ 540 kOe.<br />

0.0<br />

-0.1<br />

-0.2<br />

-0.2<br />

-0.3<br />

-0.4<br />

-0.5<br />

SrRuO 3 M 0 ∝ (T C -T) β<br />

β = 0.32<br />

T far from T C<br />

β = 0.39<br />

T near T C<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6<br />

Log(T -T)<br />

C<br />

Average β = 0.36<br />

T C = 163.2(1) K<br />

0.0<br />

125 130 135 140 145 150 155 160 165<br />

Temperature (K)<br />

Figure A- 3. Zero field magnetization M 0 of SrRuO 3 single<br />

crystal along easy [110] direction. Solid line shows the fit to<br />

M 0 (T) ∝ (1 - T/T C ) β with β = 0.36. Inset showing the same data<br />

on a log plot. The critical exponent β appears to change from<br />

Heisenberg-like β = 0.39 near T C to Ising-like β = 0.32 as T<br />

decreases.<br />

In the easy [110] direction, M vs. H curves were measured for various tem-<br />

peratures and are shown in Figure A- 2 as M 2 vs. H/M (Arrott plot). The iso-<br />

therms below T C should be approximately linear and intersect H/M = 0 at M 0 .


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

1/χ 0 (emu -1 mol Gauss)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Log(1/χ 0 )<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

3.0<br />

SrRuO 3 Crystal Inverse Susceptibility<br />

- γ<br />

χ ∝ (T-T )<br />

0 C<br />

γ = 1.17<br />

0 0.5 1 1.5<br />

Log(T-T )<br />

C<br />

0<br />

150 200 250 300 350<br />

Temperature (K)<br />

0<br />

160 165 170 175 180 185 190 195<br />

Temperature (K)<br />

From these M 0 (T), shown in Figure A- 3, the critical exponent β ≈ 0.36 and<br />

T C = 163.2 ± 0.2 K can be estimated by fitting M 0 (T) ∝ (1 - T/T C ) β in the critical<br />

region. The fit with a single value for β = 0.36 is poor considering the<br />

apparent precision of the data. Closer to T C , the data fit better with a larger<br />

β ≈ 0.39 which is close to that predicted in the 3-d Heisenberg model (β = 0.38).<br />

Farther from T C , the exponent is smaller, β ≈ 0.32, near that predicted by the<br />

Ising model (β = 0.33). It is possible that this is due to a crossover from<br />

Heisenberg to Ising behavior at T decreases from T C . A similar model has<br />

been proposed to explain measurements of thin film SrRuO 3 [73].<br />

The T > T C isotherms of Figure A- 2 should intersect M 2 = 0 at<br />

1/χ(T, H = 0) = 1/χ 0 . The critical exponent γ is then found from 1/χ 0 ∝<br />

1/ χ 0 (emu -1 mol Gauss)<br />

200<br />

150<br />

100<br />

50<br />

Pellet<br />

Single Crystal<br />

Figure A- 4. Zero field inverse susceptibility 1/χ 0 of SrRuO 3<br />

single crystal along easy [110] direction. Solid line shows the<br />

fit to 1/χ 0 (T) ∝ (1 - T/T C ) γ with γ = 1.17 and T C = 163.2 K. The<br />

inset shows the same data on a log plot.<br />

159


160 Appendix A<br />

(T/T C - 1) γ . From these data (Figure A- 4) γ = 1.17 ± 0.02 is estimated. The plot<br />

of 1/χ vs. T should be close to linear for high temperatures.<br />

2 )<br />

M 2 × |T/T C - 1| -2 β (µ B<br />

6.0<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

SrRuO 3 Crystal Scaled Arrott Plot<br />

T < T C<br />

T > T C<br />

0.0<br />

0 500 1000 1500 2000 2500<br />

H/M × |T/T - 1|<br />

C -γ (kOe/µ )<br />

B<br />

The M(H, T) data can be replotted using the scaling hypothesis [90] with<br />

β = 0.36 and γ = 1.17. According to the hypothesis, the magnetic equation of<br />

state in the critical region depends only on the scaled variables H/|T C /T - 1| β+γ<br />

and M/|T C /T - 1| β . A plot of the scaled M 2 and scaled H/M, shown in<br />

Figure A- 5, will then have only two curves: one branch for the T < T C data<br />

and another for T > T C . Not all the curves fit on a single line, this is due to<br />

the apparent change in β as discussed above.<br />

β = 0.36<br />

γ = 1.17<br />

T C =163.2K<br />

Figure A- 5. Scaled Arrott Plot of SrRuO 3 single crystal along<br />

easy [110] direction with β = 0.36 and γ =1.17. Symbols are the<br />

same as those used in Figure A- 2.<br />

The critical isotherm (T = T C ) should obey the relation M δ ∝ H, where<br />

according to the scaling relation δ = γ/β + 1. From the previously measured<br />

values of β and γ, the exponent δ is therefore expected to be 4.3 ± 0.5. This is<br />

in good agreement with the measured value (Figure A- 2) δ = 4.2 ± 0.2.


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

Magnetization (µ B /Ru)<br />

The temperature dependence of M along the easy direction is shown in<br />

Figure A- 6 for various applied fields. The demagnetization field of 4πNM ≈<br />

800 Oe should be subtracted from the applied field to give the internal field.<br />

Between 5 K and 100 K, and for internal fields from 0.2 kOe to 68 kOe the<br />

magnetization is well approximated by M = M 0 (1 - (T/Θ 2 ) 2 ) where M 0 and Θ 2<br />

are fitting parameters. Other possible analyses are discussed below.<br />

Discussion<br />

1.65<br />

1.60<br />

1.55<br />

1.50<br />

1.45<br />

1.40<br />

1.35<br />

M (µ B /Ru)<br />

SrRuO 3 Single Crystal along easy direction [110]<br />

1.7<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

0 2000 4000 6000 8000 10000<br />

T<br />

1.30<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Temperature (K)<br />

2 (K 2 )<br />

Figure A- 6. Magnetization as a function of temperature of<br />

SrRuO 3 single crystal along easy [110] direction. Inset shows<br />

the approximate T 2 dependence of the magnetization.<br />

The low temperature magnetization data (Figure A- 1) of bulk SrRuO 3 can<br />

be explained with 〈110〉 easy directions and a large cubic, but very little<br />

uniaxial, magnetocrystalline anisotropy. The low coercivity of the crystals<br />

(~10 Oe), compared to the 3 kOe coercivity found in polycrystalline samples,<br />

7T<br />

6T<br />

5T<br />

4T<br />

3T<br />

2T<br />

1T<br />

2kG<br />

1kG<br />

161


162 Appendix A<br />

provides highly reversible and square hysteresis curves. There is no<br />

indication of a magnetic multi-domain structure [183].<br />

The cubic magnetocrystalline anisotropy in SrRuO 3 is very large, as<br />

reported previously [181, 183]. Typical values of K 1 for cubic 3-d ferromagnets<br />

are 100 times smaller than that found for SrRuO 3 . Such large values of K 1<br />

usually refer to uniaxial anisotropy, for instance in hexagonal materials,<br />

which is a lower order effect (K 1 refers to the first non-zero anisotropy<br />

constant). This large anisotropy probably results from the strong spin-orbit<br />

coupling of the heavy Ru atom, which also gives SrRuO 3 a strong Kerr effect<br />

[186].<br />

From measurements along different 〈110〉 directions, no indication of<br />

uniaxial magnetocrystalline anisotropy [182] is found, although it is allowed<br />

from the orthorhombic symmetry. This might be expected since the<br />

crystallographic unit cell lengths [177] vary by only 0.03%, and the angles by<br />

0.4% from the perfect cubic ones. The distortion from cubic is primarily due<br />

to a rotation of the RuO 6 octahedra, which alters the symmetry much more<br />

than the shape of the unit cell [177]. In the crystals reported here, the<br />

orthorhombic cell was confirmed using TEM. Because the unit cell is only<br />

slightly distorted, the few reports claiming cubic [184] or tetragonal [182]<br />

crystallographic symmetry without supporting evidence, should be<br />

reevaluated in this context. Significant uniaxial anisotropy reported in thin<br />

films of SrRuO 3 [187] may result from growth induced anisotropy, as was<br />

found in films of magnetic garnets used in magnetic bubble technology [188].<br />

Due to the large magnetocrystalline anisotropy, the saturation moment<br />

of SrRuO 3 is difficult to measure. In directions other than 〈110〉 the magnetic<br />

moment does not saturate even in fields of several 10 kOe. The strain and<br />

small particle size of a polycrystalline sample apparently makes it even more<br />

difficult to saturate than the hard direction in a single crystal (Figure A- 1).<br />

Polycrystalline SrRuO 3 has a remnant magnetization (H = 0) of M S ≈ 0.85 µ B ,


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

which increases non-linearly past 1.55 µ B (H = 125 kOe) [178]. Clearly a<br />

magnetically soft single crystal with H along the easy direction is needed to<br />

measure M S . In a previous experiment [181] on single crystal SrRuO 3 with a<br />

square hysteresis loop, M S = 1.1 µ B at H = 0 was reported. However, that value<br />

of M S is clearly too small since by 17 kOe it is smaller than the value for a<br />

polycrystalline sample [181]. Our measured value of M S = 1.6 µ B (H = 0) is<br />

consistent with the high field polycrystalline results.<br />

The magnetic critical exponents measured here are in the range typically<br />

seen in large moment ferromagnets and expected theoretically for 3-<br />

dimensional ferromagnets. Experimental values for the critical exponent β in<br />

Fe, Ni and YIG [92] are 0.37 ± 0.02, which are near the theoretical values (Ising<br />

β = 0.33, Heisenberg β = 0.36). The related metallic ferromagnets La 0.5 Sr 0.5 CoO 3<br />

[92] has β = 0.361. The apparent decrease in β as T decreases from T C , may be<br />

due to the large magnetocrystalline anisotropy. A similar effect has been seen<br />

in thin film SrRuO 3 [187], where it is suggested that the magnetocrystalline<br />

anisotropy induces a crossover from Heisenberg to Ising behavior. The weak,<br />

itinerant electron ferromagnet ZrZn 2 has mean field critical exponents β = 0.5<br />

[93]. The most prominent theory on itinerant electron ferromagnetism by<br />

4/3 4/3<br />

Moriya [84, 85] predicts a TC - T dependence on the magnetization, which<br />

is essentially β = 1.<br />

The critical exponents γ and δ are also typical for large moment<br />

ferromagnets (Fe, Ni and YIG [92] γ = 1.2 ± 0.2) as opposed to those found for<br />

the weak, itinerant electron ferromagnet ZrZn 2 which has mean field critical<br />

exponents γ = 1.0 and δ = 3 [93]. The three dimensional Ising and Heisenberg<br />

models predict γ = 1.24, δ = 4.8 and γ = 1.39, δ = 4.8 respectively. The critical<br />

exponents measured here also obey the scaling relation δ = 1 + γ/β.<br />

163


164 Appendix A<br />

1/χ 0 (emu -1 mol Gauss)<br />

200<br />

150<br />

100<br />

50<br />

SrRuO 3 Polycrystal Inverse Susceptibility<br />

Pellet<br />

Single Crystal<br />

0<br />

150 200 250<br />

Temperature (K)<br />

300 350<br />

Figure A- 7. Inverse magnetic susceptibility (1/χ = M/H) at H<br />

= 10 kOe of polycrystalline SrRuO 3 compared to the single<br />

crystal data from Figure A- 4. The solid line is the straightline<br />

fit with T C = 165K which demonstrates the slightly<br />

positive curvature of the data.<br />

A positive curvature persists in the plot of 1/χ vs. T (Figure A- 7) even at<br />

higher temperature where it should be linear for a Curie-Weiss ferromagnet.<br />

Such a curvature can be caused by a temperature independent term in the<br />

susceptibility χ = χ Const + C/(T - Θ), where C/(T - Θ) is the Curie-Weiss<br />

susceptibility and is always positive (T > Θ). A χ Const < 0 of about -4 × 10 -4 emu<br />

G -1 mol -1 will provide the observed curvature, and has been independently<br />

observed elsewhere [175]. The temperature independent term χ Const , should<br />

contain a positive contribution due to Pauli paramagnetism. This can be<br />

estimated from measurements of the linear term of the specific heat [176],<br />

giving χ Pauli ≈ +4 × 10 -4 emu G -1 mol -1 . The Landau diamagnetism should be


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

negative and for simple band structures is smaller than χ Pauli (for free electrons<br />

χ Landau = -χ Pauli /3). The core diamagnetism can be estimated from tables of<br />

experimental values [76] giving χ Core = -0.7 × 10 -4 emu G -1 mol -1 . The sum of<br />

these theoretical estimates χ Const = χ Pauli + χ Landau + χ Core is, however, still<br />

positive while the experimental value appears to be negative. In the critical<br />

region, 1/χ ∝ (T/T C - 1) γ with γ = 1.17, provides a positive curvature in the plot<br />

of 1/χ vs. T. Since the mean field exponent γ = 1 is expected to be valid far<br />

from T C , some other mechanism must provide the effective γ > 1 observed at<br />

these higher temperatures.<br />

Magnetic excitations which become thermally induced as the temperature<br />

is raised above T = 0 reduce the magnetization from the ground state value.<br />

The exponential decrease predicted by the mean field model has some<br />

qualitative value but is never in good agreement with experiment. Collective<br />

spin wave excitations and single particle (Stoner) excitations both decrease the<br />

magnetization according to a power law M = M S (1 - (T/Θ n ) n ) which is in accord<br />

with experiments where n ≈ 2 ± 1 and Θ n is of the order T C .<br />

The limiting low T, H = 0 behavior of collective, spin wave excitations<br />

(section 3.2.2.2.8) predicts n = 3/2 and for SrRuO 3 , Θ 3/2 ≈ 2.42T C = 400 K. The<br />

Stoner theory of single (k-space) particle excitations (section 3.2.2.2.2) predicts<br />

n = 2 and Θ 2 ≈ 1.41 T C (= 230 K for SrRuO 3 ). As the temperature and magnetic<br />

field is raised, these models require further corrections. For example, higher<br />

order corrections such as an additional n = 5/2 term in predicted in the spin<br />

wave theory. For H ≠ 0, the low temperature spin-wave excitations become<br />

quenched which can have the effect of increasing the average value on n [97].<br />

The generalized model of spin fluctuations in ferromagnets ([84] section<br />

165


166 Appendix A<br />

T 3/2 Coefficient (10 -6 /K 3/2 )<br />

200<br />

150<br />

100<br />

50<br />

0<br />

A parameter in fit to M = M S (1 - AT 3/2 - BT 2 )<br />

-50<br />

0 25 50 75 100 125<br />

Maximum Temperature of Fitting Range<br />

150<br />

3.2.2.2.3) includes both types of interacting magnetic excitations and predicts<br />

n = 3/2 for H = 0, T = 0 but also n ≈ 2 in calculations [86].<br />

T 2 Coefficient (10 -6 /K 2 )<br />

25<br />

20<br />

15<br />

Correlation of A and B<br />

10<br />

-20 0 20 40 60 80<br />

T 3/2 Coefficient (10 -6 /K 3/2 )<br />

Figure A- 8. Variation of the T 3/2 parameter in fitting the<br />

magnetization data of single crystal SrRuO 3 to M = M S (1 -<br />

AT 3/2 - BT 2 ) as the fitting range is increased. The upper inset<br />

shows the correlation of the A and B parameters. In the<br />

region where A is relatively stable (around T max = 60 K), A<br />

decreases as T max is lowered. The symbols are the same as<br />

those used in Figure A- 9.<br />

Experimental results on metallic ferromagnets with substantial saturation<br />

moments such as Fe, Ni [88, 189-192] and La 0.67 Sr 0.33 MnO 3 ([97], section 4.1.1)<br />

can display n = 3/2 consistent with the spin wave stiffness determined by<br />

neutron diffraction if significant corrections are included and only low<br />

temperature data is analyzed. In contrast “weak” itinerant-electron


T 2 Coefficient (10 -6 /K 2 )<br />

Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

50<br />

40<br />

30<br />

20<br />

10<br />

-10<br />

0 25 50 75 100 125 150<br />

Maximum Temperature of Fitting Range<br />

ferromagnets such as ZrZn 2 , Ni 3 Al and Sc 3 In, show n = 2 [86, 193-195] over a<br />

wide temperature range.<br />

0<br />

As shown above (Figure A- 6) the magnetization of SrRuO 3 single crystal<br />

can be well described by M = M S (1 - (T/Θ n ) n ) with n = 2 in the temperature<br />

range of reliable measurement, 5 K to 100 K. A fit with n = 3/2 in this<br />

temperature range is unsatisfactory.<br />

B parameter in fit to M = M 0 (1 - AT 3/2 - BT 2 )<br />

7T T 2 term<br />

6T T 2 term<br />

5T T 2 term<br />

4T T 2 term<br />

3T T 2 term<br />

2T T 2 term<br />

1T T 2 term<br />

2kOe T 2 term<br />

1kOe T 2 term<br />

Figure A- 9. Variation of the T 2 parameter in fitting the<br />

magnetization data of single crystal SrRuO 3 to M = M S (1 -<br />

AT 3/2 - BT 2 ) as the fitting range is increased. In the region<br />

where B is relatively stable (around T max = 60 K), B increases as<br />

T max is lowered.<br />

If one assumes n = 2 arises from Stoner excitations and collective<br />

excitations result in n = 3/2, then one may expect the variation of<br />

167


168 Appendix A<br />

magnetization with temperature to fit with a combination of the n = 2 and<br />

n = 3/2 terms, M = M S (1 - AT 3/2 - BT 2 ) [150] at low temperatures. Since an<br />

additional parameter is included, an improvement of the fit does not<br />

necessarily prove the significance of the added parameter. Furthermore, the<br />

functions T 3/2 and T 2 are very similar making the fitting parameters highly<br />

correlated.<br />

The magnetization data for single crystal SrRuO 3 were fit to a polynomial<br />

expression M = M S (1 - AT 3/2 - BT 2 ) from T = 2 K to T = T max . As expected, the A<br />

and B parameters are highly correlated (Figure A- 8) with ∆B ≈ 2.2 × 10 -5 K -2 -<br />

0.2 K -1/2 ∆A. The parameters A and B can then be plotted as a function of T max<br />

(Figure A- 8 and Figure A - 9). For very low T max (< 30 K) the fit is unstable<br />

since the difference in the magnetization at T


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

T 2 Coefficient (10 -6 /K 2 )<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

B parameter in fit to M = M 0 (1 - BT 2 )<br />

-10<br />

0 25 50 75 100 125 150<br />

Maximum Temperature of Fitting Range<br />

In fitting to the data on SrRuO 3 single crystal the B parameter (Figure A-9)<br />

is relatively stable and increases as lower temperature fitting ranges (T max ) are<br />

used. The A parameter (Figure A- 8) is not as stable and appears to decrease as<br />

T max is lowered. If this trend were to continue, then the T 3/2 contribution<br />

would be very small at low temperatures. The removal of the AT 3/2 term<br />

further stabilizes the B parameter (Figure A- 10), and makes B more<br />

independent of T max than with AT 3/2 included. Thus, from this analysis of the<br />

magnetization data, there is little evidence for a large AT 3/2 contribution to<br />

the magnetization. This is in contrast with measurements on thin film [186,<br />

187] which show T 3/2 dominating over T 2 .<br />

Θ 2 (K)<br />

240<br />

235<br />

230<br />

233K = √2T<br />

C<br />

expected from<br />

Θ (T = 60K)<br />

2 max<br />

225 simple Stoner Theory<br />

220<br />

Θ = B 2 1/2<br />

215<br />

210<br />

0 1 2 3 4 5 6 7<br />

Internal Field (T)<br />

Figure A- 10. Variation of the T 2 parameter in fitting the<br />

magnetization data of single crystal SrRuO 3 to M = M S (1 - BT 2 )<br />

as the fitting range is increased. The parameter B for this fit is<br />

more stable and constant than that shown in Figure A- 8.<br />

Inset, variation of Θ 2 in a magnetic field.<br />

The two corrections to the spin wave T 3/2 theory mentioned above (finite<br />

magnetic field and higher order T (2n+1)/2 terms) may not adequately explain the<br />

data presented here. The applied magnetic field will suppress spin wave<br />

169


170 Appendix A<br />

excitations for T < gµ B H/k B , resulting in an effective n > 3/2. If this were a<br />

significant effect, the T 3/2 fitting parameter A should decrease as H is increased<br />

particularly as T max approaches 0 K. Such a systematic decrease in A is not<br />

obviously apparent (Figure A- 8). One cannot however, exclude the<br />

demagnetization or anisotropy field which are present even when the applied<br />

H = 0.<br />

AT 3/2 Coefficient (10 -6 /K 3/2 )<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

M = 1 - (T/400 K) 3/2 - (T/600 K) 5/2<br />

fit with M = M (1 - AT<br />

S 3/2 - BT2 )<br />

A<br />

B<br />

0<br />

0.0<br />

0 20 40 60 80 100<br />

Maximum Temperature (K) of Fitting Range<br />

Figure A- 11. Variation of A and B fitting parameters in the<br />

hypothetical case where the true magnetization is given by T 3/2<br />

and T 5/2 terms.<br />

If the true magnetization were a sum of T 3/2 and T 5/2 terms, M/M S =<br />

(1 - (T/400 K) 3/2 - (T/600 K) 5/2 ) for instance, the data would fit well to M/M S =<br />

(1 - AT 3/2 - BT 2 ). The fitting parameters A and B as a function of T max for this<br />

example are shown in Figure A- 11. The A parameter is relatively constant,<br />

increasing slightly as T max approaches 0 K. In the limit T max approaches 0 K,<br />

A = 125 × 10 -6 K -3/2 or Θ 3/2 = 400.3 K, which is very close to the true value of<br />

400 K. The B fitting parameter is not nearly as stable and decreases as T max<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

BT 2 Coefficient (10 -6 /K 2 )


Critical Behavior and Anisotropy in Single Crystal SrRuO 3<br />

approaches 0 K. Since the exponent of the BT 2 fit (2) is smaller than 5/2,<br />

which is the exponent used in the example, a larger B is needed to fit the data<br />

at higher temperatures. Thus, a dramatically decreasing B as T max approaches<br />

0 K, is indicative of a fitting exponent (2 in this example) smaller than the<br />

true exponent (5/2 in this example).<br />

The data on SrRuO 3 shows the opposite effect: B is relatively constant<br />

while A decreases as T max approaches 0 K. This decrease of A indicates that<br />

the exponent 3/2 is smaller than the true exponent, and therefore, the<br />

addition of higher order terms to the spin wave form of the magnetization<br />

will not fully explain the data on SrRuO 3 .<br />

Positive B (and A) parameters can be converted to the Θ n values used<br />

above via Θ 2 = B -1/2 and Θ 3/2 = A -2/3 . The Θ 2 values extracted from fits<br />

described above are shown as a function of the internal magnetic field in<br />

Figure A- 10. The values of Θ 2 (H) extrapolated to T max = 0 are 20 K less than<br />

the corresponding Θ 2 (H) for T max = 60 K. The single crystal Θ 2 parameters<br />

increase only slightly as the field is increased, about 2.6 K/Tesla. These values<br />

for Θ 2 are consistent with measurements of the field dependent heat capacity<br />

of polycrystalline SrRuO 3 (Figure B- 3) via the Maxwell relation<br />

⎛ ∂CH ⎞<br />

⎝ ∂H ⎠ T<br />

= T ∂ 2 M<br />

∂T 2<br />

⎛ ⎞<br />

⎜ ⎟ .<br />

⎝ ⎠ H<br />

Conclusion<br />

Various magnetic properties of single crystal SrRuO 3 along the magnetic<br />

easy direction have been measured. The saturation moment, 1.61µ B /Ru, is<br />

larger than that reported previously for single crystal material but in accord<br />

with experiments on polycrystalline material and theoretical calculations.<br />

The crystals show extremely large cubic magnetocrystalline anisotropy while<br />

no substantial uniaxial magnetocrystalline anisotropy. The critical exponent<br />

171


172 Appendix A<br />

β changes from Heisenberg like 0.39 near T C to Ising like 0.32 further from T C .<br />

The susceptibility exponent γ = 1.17 persists to 2 T C , where it is expected to<br />

decrease to 1. The magnetization from 5 K to 100 K is well described by<br />

M = M S (1 - (T/Θ n ) n ) where n ≈ 2. The addition of a T 3/2 term is not obviously<br />

significant.


Appendix B. Magnetic Excitations and Specific Heat in<br />

SrRuO 3<br />

Low temperature magnetic excitations can be probed not only by<br />

measurements of the magnetization but also by the heat capacity. As will be<br />

discussed below, the two measurements are related via a Maxwell relation.<br />

SrRuO 3 polycrystalline samples were prepared by repeated heating of a R u<br />

and SrCO 3 mixture at 1260°C (seciton 2.1.1). Heat capacity data were taken<br />

using the relaxation method [102] (section 3.3). Constant field data were taken<br />

on a zero field cooled sample at 0 and 8 T, and in zero field after the<br />

application of an 8 T field. In addition, data at temperatures between 4 K and<br />

5 K were taken at several intermediate fields.<br />

Specific heat data taken in 0 T and 8 T applied magnetic field are shown in<br />

Figure B- 1. The zero field cooled and zero field remnant data are equal<br />

within experimental error. Between 3 K and 13 K both the 0 T and 8 T data<br />

appear to be well described by a straight line on this plot of c/T vs. T 2 ,<br />

suggesting that the specific heat can be described by γT+βT 3 . The parameter<br />

values from fitting the data in zero field are γ = 37 mJ mol -1 K -2 and<br />

β = 0.19 mJ mol -1 K -4 . This value of β corresponds to a Debye temperature of<br />

371 K and is in good agreement with previous zero-field measurements [176,<br />

196], while γ is about 20% larger than that found previously. The value of γ in<br />

zero field, 37 mJ mol -1 K -2 , is 4.6 times that expected of the bare density of states<br />

given by spin resolved band structure calculations [176]. The effect of<br />

demagnetization fields should be negligible since the correction to the specific<br />

heat is approximately proportional to T 3 but only about 0.1% of the measured<br />

value of β [197, 198].<br />

173


174 Appendix B<br />

c P,H /T (mJ/mol·K 2 )<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

SrRuO 3 Specific Heat<br />

0 T zfc<br />

0 T rem<br />

8 T<br />

-5<br />

0 50 100 150 200 250 300 350<br />

Temperature 2 (K 2 )<br />

30<br />

0 50 100 150 200 250 300 350<br />

Temperature 2 (K 2 )<br />

Above 13 K and below 3 K the data in Figure B- 1 deviate from a straight<br />

line. The deviation above 13K can be fit by an expression for the specific heat<br />

of a thermally activated Einstein (optical) phonon mode, with an Einstein<br />

temperature of Θ Ε = 140 K. Both β and Θ Ε are field independent within<br />

experimental error. The temperature range below 3 K is not large enough to<br />

draw conclusion about the functional form of the low-temperature upturn.<br />

A field independent term proportional to T 3 ln(T) is expected from the non-<br />

spinwave spin fluctuations [91, 194, 199]. A Shottky anomaly due to free<br />

∆c P,H /T (mJ/mol·K 2 )<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

c P,H /T (rem) - c P,H /T (zfc)<br />

c P,H /T (8T) - c P,H /T (zfc)<br />

Figure B- 1 Heat capacity of SrRuO 3 cooled in zero field (zfc),<br />

in an 8 T magnetic field, and in zero field after being<br />

magnetized (rem). Inset, difference between the heat capacity<br />

measured after cooling in zero field with that in 8 T and the<br />

remnant magnetized state.


c P,H /T (mJ/mol·K)<br />

65<br />

60<br />

55<br />

50<br />

45<br />

40<br />

Specific Heat of SrRuO 3<br />

SrRuO 3 Specific Heat in Zero Field<br />

c P,H (T) = γT + βT 3 +<br />

2mJ/(mol·K 5/2 ) T 3/2<br />

c P,H (T) = γT + βT 3<br />

H=0 Remnant<br />

H=0 Zero Field Cooled<br />

35<br />

0 25 50 75 100 125 150<br />

Temperature 2 (K 2 )<br />

Figure B- 2. Zero field heat capacity data fit with two free<br />

parameters, γ and β. Solid line, including the T 3/2 contribution<br />

expected theoretically for spin waves (see text). Dashed line,<br />

without any T 3/2 contribution.<br />

paramagnetic spins will give a low temperature upturn but should also be<br />

noticeable as Curie paramagnetism in magnetization measurements.<br />

Adding a positive T 3/2 term to the model for the specific heat, as expected<br />

for pure spin wave excitations (3.2.2.2.8) [95], does not improve the fit (Figure<br />

B- 2). Assuming 6 nearest neighbors, spin 1, and T C = 163 to determine J, this<br />

value is A = 2mJ mol -1 K -5/2 . Previous magnetization measurements on thin<br />

films suggest that this term is correct within a factor of 2 [186]. Using a<br />

functional form of γT + AT 3/2 + βT 3 , the coefficient A is less than 1.2 (0 T zfc),<br />

2.5 (8 T), and 0.6 (0 T rem) mJ mol -1 K -5/2 for each of the three data sets i n<br />

175


176 Appendix B<br />

Figure 3 at the 90% statistical confidence level. The T 3/2 term in Fe and Ni is<br />

quite small but has nevertheless been extracted from the heat capacity data<br />

[200-202].<br />

γ(H) (mJ/mol-K 2 )<br />

Field Dependence of the Density of States in SrRuO 3<br />

38<br />

37<br />

36<br />

35<br />

34<br />

34<br />

33<br />

32<br />

Predicted from M vs. T<br />

Single Crystal Data<br />

0 2 4 6 8<br />

Magnetic Field (Tesla)<br />

Figure B- 3. The linear term of the heat capacity γ as a<br />

function of magnetic field. Each circle is from a single c P,H<br />

datum between 4.3 and 5 K with phonons subtracted: γ(H) =<br />

(c P,H (T) - βT 3 )/T with β = 0.191 mJ/mol·K 4 . γ(H) for each square<br />

was determined by fitting 15-20 data points between 6 and 12 K<br />

to: c P,H (T) = γT + βT 3 . The triangles are calculated from the<br />

magnetization data of a single crystal.<br />

The linear-T term, γ, is significantly depressed in an 8 T magnetic field.<br />

The inset to Figure B- 1 shows the difference between the zero-field cooled<br />

data in zero field and each of the other two data sets, plotted as ∆c/T. The data


Specific Heat of SrRuO 3<br />

taken in 0 T after application of an 8 T field (circles) shows that the zero-field<br />

data are not hysteretic. On this plot of ∆c/T, the 8 T data (diamonds) show<br />

that the only observed change in the specific heat is a decrease in the linear-T<br />

term.<br />

To further understand the field dependence of γ, specific heat points<br />

between 4 K and 5 K at several intermediate fields were also taken. These<br />

data are shown in Figure B- 3 (open circles) as γ(H) = (c(T,H) - βT 3 )/T using the<br />

value of β measured at 0 T and 8 T. On the same plot, the open squares<br />

indicate the fitted value of γ from the three data sets in Figure B- 1.<br />

Figure B- 3 shows that the coefficient of the linear-T term in the specific heat<br />

decreases approximately linearly with applied magnetic field, H.<br />

To summarize, the specific heat between 3 and 13 K can be described by<br />

c(T,H) = (γ - γ ´H)T+βT 3 .<br />

Although the magnetization of polycrystalline SrRuO 3 is irreversible, the<br />

measured heat capacity is not. Therefore, the heat capacity originates from<br />

processes reversible with respect to a magnetic field, showing that the entropy<br />

associated with random domains is negligible. In this sense, the heat capacity<br />

is a better thermodynamic measure of the low-lying excitations. The<br />

magnetic field dependence of the heat capacity and the magnetization are<br />

related by the following Maxwell relation:<br />

⎛ ∂CH ⎞<br />

⎝ ∂H ⎠ T<br />

= T ∂ 2 M<br />

∂T 2<br />

⎛ ⎞<br />

⎜ ⎟<br />

⎝ ⎠ H<br />

(1)<br />

According to equation 1, the first derivative of the heat capacity with<br />

respect to field gives information about the second derivative of the<br />

magnetization with respect to temperature. This relation is generally valid<br />

for an isotropic system as long as there are no irreversible processes involved<br />

and no PV work is done. Equation 1 is actually three equations since H and M<br />

177


178 Appendix B<br />

are vectors. Thus for an ideal, simple experiment, the sample would be very<br />

soft magnetically (no magnetic hysteresis) and the magnetic field would be<br />

aligned in the magnetic easy direction to keep M and H parallel. If possible,<br />

the magnetic and heat capacity experiments should be done on the same<br />

sample in the same direction; although, in principle different samples should<br />

give the same result since both M and C P in a reversible sample should not<br />

depend on the microstructure. In order to use equation 1 to relate the<br />

reversible magnetization from the measured specific heat, it is assumed that<br />

effects due to the field dependence of the magnetocrystalline anisotropy are<br />

insignificant.<br />

The simplest spin wave model (section 3.2.2.2.8) predicts a T 3/2 dependence<br />

of the magnetization at low temperatures:<br />

MHT ( , ) T<br />

≈1− MH ( , 0)<br />

3 ( H)<br />

⎛ ⎞<br />

⎜ ⎟<br />

⎝ Θ ⎠<br />

with Θ 3/2 ≈ 2.42T C = 400K [95, 186] for SrRuO 3 . Corrections due to H ≠ 0 and<br />

T > 0 can also be calculated [95]. The Stoner/Wohlfarth theory of magnetism<br />

(section 3.2.2.2.2) [81] predicts a somewhat different temperature dependence:<br />

MHT ( , ) T<br />

≈1− MH ( , 0)<br />

( H)<br />

⎛ ⎞<br />

⎜ ⎟<br />

⎝ Θ ⎠<br />

(Θ 2 ≈ √2T C = 231K for SrRuO 3 ) due to single (k-space) particle excitations.<br />

The Maxwell relation together with the observed magnetic field<br />

dependence of the specific heat above 4 K, γ ´HT, implies that the second<br />

2<br />

⎛ ∂ M⎞<br />

derivative of the magnetization with respect to temperature, ⎜ 2 ⎟ , is a<br />

⎝ ∂T<br />

⎠ H<br />

constant. This result is consistent with the dominant T 2 temperature<br />

dependence of the reversible discussed above.<br />

In order to make quantitative comparisons between the specific heat data<br />

and the magnetization data, equation 1 can be integrated with respect to H<br />

2<br />

2<br />

3 2<br />

2<br />

(2)<br />

(3)


Specific Heat of SrRuO 3<br />

from H = 0 to H = ∆H. Since magnetization data is not available at all fields,<br />

the integral is approximated to a finite sum with step size dH:<br />

∆H/<br />

dH−<br />

∆CH(<br />

T)<br />

dH M( n ⋅ dH, T)<br />

≈ ∑ T∆H ∆Hn=<br />

T<br />

∂<br />

1 2<br />

(4)<br />

2<br />

0 ∂<br />

where ∆CH (T)=CH (T,H=∆H)-CH (T,H=0). The two simplest physically plausible<br />

temperature dependencies of the reversible magnetization, T 2 (equation 3) or<br />

T 3/2 (equation 2), each correspond to a simple temperature dependence of<br />

∆CH<br />

T<br />

T∆H ( ) :<br />

H/ dH 1<br />

MHT ( , ) T CH( T)<br />

2dH Mn ( dH,<br />

0)<br />

≈1− Const.<br />

0<br />

2<br />

MH ( , 0)<br />

( H)<br />

T H H ( n dH)<br />

⎛<br />

∆ −<br />

⎞ ∆ − ⋅<br />

⎜ ⎟ ⇒ = ∑<br />

= <<br />

⎝ Θ ⎠ ∆ ∆ Θ ⋅<br />

2<br />

2<br />

H/ dH 1<br />

MHT ( , ) T CH( T)<br />

3dH<br />

Mn ( dH,<br />

0) 1<br />

≈1− 3<br />

2<br />

MH ( , 0)<br />

3 ( H)<br />

T H 4 T H n 0 3 ( n dH) T<br />

⎛ ⎞<br />

− ⋅<br />

⎜ ⎟ ⇒ = ∝<br />

⎝ ⎠<br />

⋅<br />

−<br />

∆ −<br />

∆<br />

∑<br />

Θ<br />

∆ ∆ = Θ<br />

2<br />

3 2<br />

Since the ∆cPH<br />

, T<br />

T∆H ( ) data of Figure B- 1 are constant in temperature, these<br />

data can be fit to a T 2 dependence of the magnetization, but not by T 3/2 alone.<br />

Using the parameters from the fits to the magnetization data of single crystal<br />

SrRuO3 (Appendix A) gives quantitative values for the constant ∆cPH<br />

, T<br />

T∆H ( ) =<br />

-0.42 mJ mol -1 K -2 T -1 (about 20% less than that measured by ∆cPH<br />

, T<br />

T∆H ( ) ). A field<br />

independent Θ2 (H) is consistent with γ decreasing linearly with magnetic<br />

field. Since the two measurements are on different types of samples with<br />

different orientations; nevertheless this result illustrates the potential for this<br />

method.<br />

The failure to observe a T 3/2 dependence does not necessarily eliminate the<br />

possibility of spin wave excitations in SrRuO 3 ; the data could be easily<br />

described within the spin wave theory (section 3.2.2.2.8) if additional<br />

n=<br />

0<br />

2<br />

2<br />

179<br />

(5)<br />

(6)


180 Appendix B<br />

corrections are included. Using these corrections, the predicted<br />

magnetization fit to T n will necessarily have n ≈ 2 to fit the data.<br />

At low temperatures and finite fields, a gap is expected in the spin wave<br />

dispersion [95], predicting an exponential temperature dependence of the<br />

magnetization at the lowest temperatures. Combined with the spin wave n =<br />

3/2, H = 0 form, this effectively increases n [97]. The inclusion of this gap<br />

affects the predicted ∆cPH<br />

, T<br />

T∆H ( ) primarily at low temperatures, as is shown in<br />

(Figure 3-11).<br />

The linear term in the specific heat γ is usually interpreted in terms of the<br />

(nonmagnetic) electronic contribution to the specific heat. Spin fluctuations<br />

are predicted to enhance γ in ferromagnetic and nearly ferromagnetic metals<br />

[84]. For γ = γ ele (1 + λ e-p + λ mag ) where γ ele is γ expected from band calculations<br />

and λ e-p ≈ 0.5 is the electron-phonon mass enhancement [176]. From these<br />

values λ mag ≈ 3.1, the electron-spin fluctuation mass enhancement, is<br />

estimated. Theoretically, in the presence of a magnetic field, spin fluctuations<br />

are suppressed and therefore γ will decrease [203, 204]. From equation 1, these<br />

spin fluctuations which cause the field dependence of γ are required to give a<br />

T 2 dependence of the magnetization, thereby unifying the two interpretations<br />

for the field dependence of γ. A decrease in the coefficient of the linear-T<br />

term, γ, has also been observed in the weak itinerant ferromagnets ZrZn 2 and<br />

Sc 3 In [199, 205] (which also show a T 2 dependence of the magnetization).<br />

In conclusion, the magnetic field dependence of the heat capacity provides<br />

a useful method for studying the low-lying excitations, and for determining<br />

the temperature dependence of the magnetization via thermodynamic<br />

relations. For a ferromagnetic substance the magnetic field dependence of the<br />

low-temperature heat capacity can be a more sensitive probe of the low-


Specific Heat of SrRuO 3<br />

energy magnetic excitations than the saturation magnetization, particularly<br />

for polycrystalline samples with large magnetocrystalline anisotropy such as<br />

SrRuO 3 .<br />

Like the magnetization measurements, the heat capacity measurements<br />

can be well described by a decrease in the magnetization proportional to T n<br />

where n ≈ 2. This can be understood by any of three mechanisms. First, if the<br />

corrections are included in the spin wave theory then an effective n > 3/2 is<br />

predicted. Second, single particle excitations are present. Third, the<br />

excitations interact (requiring a more generalized description [85]) to give an<br />

effective exponent n ≈ 2.<br />

The same phenomena can also be interpreted as a spin fluctuation<br />

enhancement of the electronic specific heat, which can be suppressed by a<br />

magnetic field.<br />

The help of K. A. Moler for the heat capacity measurements is greatly<br />

appreciated.<br />

181


References<br />

[1] G. H. Jonker and H. van Santen, Physica 16, 337 (1950).<br />

[2] H. van Santen and G. H. Jonker, Physica 16, 599 (1950).<br />

[3] E. O. Wollan and W. C. Koehler, Phys. Rev. 100, 545 (1955).<br />

[4] J. Volger, Physica 20, 49 (1954).<br />

[5] J. B. Goodenough and J. M. Longo, in Landolt-Börnstein Tabellen<br />

(Springer, Berlin, 1970), Vol. III/4a.<br />

[6] C. Zener, Phys. Rev. 82, 403 (1951).<br />

[7] P. W. Anderson and H. Hasegawa, Phys. Rev. 100, 675 (1955).<br />

[8] P.-G. deGennes, Phys. Rev. 118, 141 (1960).<br />

[9] S. Zhang, J. Appl. Phys. 79, 4542 (1996).<br />

[10] A. J. Millis, P. B. Littlewood, and B. I. Shraiman, Phys. Rev. Lett. 74, 5144<br />

(1995).<br />

[11] A. J. Millis, B. I. Shraiman, and R. Mueller, Phys. Rev. Lett. 77, 175 (1996).<br />

[12] A. J. Millis, B. I. Shraiman, and R. Mueller, Phys. Rev. B 53, 8434 (1996).<br />

[13] C. B. Alcock, R. C. Doshi, and Y. Shen, Solid State Ionics 51, 281 (1992).<br />

[14] T. Kobayashi, H. Takizawa, T. Endo, T. Sato, and M. Shimada, J. Solid<br />

State Chem. 92, 116 (1991).<br />

[15] H. Taguchi, M. Nagao, and M. Shimada, J. Solid State Chem. 97, 476<br />

(1992).<br />

[16] J. A. M. van Roosmalen, J. P. P. Huijsmans, and L. Plomp, Solid State<br />

Ionics 66, 279 (1993).<br />

[17] A. Hammouche, E. J. L. Schouler, and M. Henault, Solid State Ionics 28-<br />

30, 1205 (1988).<br />

[18] H. L. Ju, C. Kwon, Q. Li, R. L. Greene, and T. Venkatesan, Appl. Phys. Lett.<br />

65, 2108 (1994).<br />

[19] J. Cho, M. Gomi, and M. Abe, Jpn. J. Appl. Phys. 29, 1686 (1990).<br />

182


References 183<br />

[20] K. C. Chanara, T. Ohno, M. Kasai, and Y. Kozono, Appl. Phys. Lett. 63,<br />

1990 (1993).<br />

[21] R. von Helmolt, J. Wecker, B. Holzapfel, L. Schultz, and K. Samwer, Phys.<br />

Rev. Lett. 71, 2331 (1993).<br />

[22] S. Jin, T. H. Tiefel, M. McCormack, R. A. Fastnacht, R. Ramesh, and L. H.<br />

Chen, Science 264, 413 (1994).<br />

[23] G. C. Xiong, Q. Li, H. L. Ju, S. N. Mao, L. Senapati, X. X. Xi, R. L. Greene,<br />

and T. Venkatesan, Appl. Phys. Lett. 66, 1427 (1995).<br />

[24] J. Z. Sun, L. Krusin-Elbaum, S. S. P. Parkin, and G. Xiao, Appl. Phys. Lett.<br />

67, 2726 (1995).<br />

[25] S. Jin, M. McCormack, T. H. Tiefel, and R. Ramesh, J. Appl. Phys. 76, 6929<br />

(1994).<br />

[26] S. Kawamata, N. Kobayashi, M. Ikebe, and Y. Muto, Physica B 148, 130<br />

(1987).<br />

[27] G. C. Xiong, Q. Li, H. L. Ju, R. L. Greene, and T. Venkatesan, Appl. Phys.<br />

Lett. 66, 1689 (1995).<br />

[28] N. Furukawa, J. Phys. Soc. Jpn. 65, 1174 (1996).<br />

[29] P. Schiffer, A. P. Ramirez, W. Bao, and S.-W. Cheong, Phys. Rev. Lett. 75,<br />

3336 (1995).<br />

[30] C. M. Varma, Phys. Rev. B 54, 7328 (1996).<br />

[31] N. Furukawa, J. Phys. Soc. Jpn. 63, 3214 (1994).<br />

[32] S. Zhang and Z. Yang, J. Appl. Phys. 79, 7398 (1996).<br />

[33] M. S. Hillery, D. Emin, and N.-L. H. Liu, Phys. Rev. B 38, 9771 (1988).<br />

[34] D. Emin, M. S. Hillery, and N.-L. H. Liu, Phys. Rev. B 35, 641 (1987).<br />

[35] A. World and K. Dwight, J. Solid State Chem. 88, 229 (1990).<br />

[36] R. J. Bouchard and J. L. Gillson, Mat. Res. Bull. 7, 873 (1972).<br />

[37] R. Czaya, Z. anorg. allgem. Chem. 375, 61 (1970).<br />

[38] G. J. Snyder and A. Simon, Angew. Chem. Int. Ed. 33, 689 (1994).<br />

[39] G. J. Snyder and A. Simon, J. Amer. Chem. Soc. 117, 1996 (1995).


184 References<br />

[40] G. J. Snyder and A. Simon, Adv. Mater. 6, 374 (1994).<br />

[41] G. J. Snyder, M. Gelabert, and F. J. DiSalvo, J. Solid State Chem. 113, 355<br />

(1994).<br />

[42] G. J. Snyder, M. E. Badding, and F. J. DiSalvo, Inorg. Chem. 31, 2107 (1992).<br />

[43] F. J. DiSalvo, Current Opinion in solid state materials science 1, 241-9<br />

(1996).<br />

[44] G. J. Snyder and A. Simon, Zeit. Kryst. 209, 384 (1994).<br />

[45] G. J. Snyder and A. Simon, Zeit. Natur. B 49, 189 (1994).<br />

[46] G. J. Snyder and A. Simon, J. Chem. Soc. (London) Dalton Trans. 1994<br />

(1994).<br />

[47] G. J. Snyder, H. Borrmann, and A. Simon, Zeit. Kryst. 209, 458 (1994).<br />

[48] G. J. Snyder and A. Simon, J. Alloys Compounds 223, 65 (1995).<br />

[49] T. M. Hayes and J. B. Boyce, in Solid State Physics, edited by H.<br />

Ehrenreich, F. Seitz and D. Turnbull (Academic, New York, 1982), Vol. 37,<br />

p. 173.<br />

[50] J. Antoszewski, D. J. Seymour, and L. Faraone, J. Electr. Mater. 24, 1255<br />

(1995).<br />

[51] F. M. Smits, The Bell System Technical Journal 1958, 711 (1958).<br />

[52] L. J. van der Pauw, Philips Res. Repts. 13, 1 (1958).<br />

[53] L. J. van der Pauw, Philips Tech. Rev. 20, 220 (1958).<br />

[54] D. K. C. MacDonald, in Handbuch der Physik, edited by S. Flügge<br />

(Springer Verlag, Berlin, 1953), Vol. 14.<br />

[55] J. M. Ziman, Electrons and Phonons (Clarendon press, Oxford, 1960).<br />

[56] N. W. Ashcroft and N. D. Mermin, Solid State Physics (Saunders, 1976).<br />

[57] H. Frölich, Adv. Phys. 3, 325 (1954).<br />

[58] D. Emin and T. Holstein, Ann. Phys. 53, 439 (1969).<br />

[59] D. Emin, Phys. Rev. B 3, 1321 (1971).<br />

[60] T. Holstein, Ann. Phys. (N.Y.) 8, 343 (1959).<br />

[61] D. Emin and N. L. H. Liu, Phys. Rev. Lett. 42, 71 (1979).


References 185<br />

[62] D. Emin and N. L. H. Liu, Phys. Rev. B 27, 4788 (1983).<br />

[63] L. Friedman and T. Holstein, Ann. Phys. (N.Y.) 21, 494 (1963).<br />

[64] T. Holstein, Ann. Phys. (N.Y.) 8, 325 (1959).<br />

[65] D. Emin, Adv. Phys. 24, 305 (1975).<br />

[66] Shewmon, Diffusion in solids (Minerals Metals and Materials Society,<br />

Warrendale, PA, 1989).<br />

[67] P. A. Cox, The Electronic Structure and Chemistry of Solids (Oxford<br />

University Press, Oxford, 1987).<br />

[68] J. B. Goodenough, Progress in Solid State Chemistry 5, Chap. 4 (1970).<br />

[69] N. F. Mott, Conduction in non-Crystalline Materials (Clarendon Press,<br />

Oxford, 1993).<br />

[70] N. F. Mott and E. A. Davis, Electronic Processes in non-Crystalline<br />

Materials (Clarendon Press, Oxford, 1979).<br />

[71] B. I. Shklovskii and A. L. Efros, Electronic Properties of Doped<br />

Semiconductors (Springer-Verlag, 1984).<br />

[72] M. I. Klinger, J. Phys. C: Solid State Phys. 9, 3955 (1976).<br />

[73] L. Klein, J. S. Dodge, C. H. Ahn, G. J. Snyder, T. H. Geballe, M. R. Beasley,<br />

and A. Kapitulnik, Phys. Rev. Lett. 77, 2774 (1996).<br />

[74] A. P. Ramirez, P. Schiffer, S.-W. Cheong, C. H. Chen, W. Bao, T. T. M.<br />

Palstra, P. L. Gammel, D. J. Bishop, and B. Zegarski, Phys. Rev. Lett. 76,<br />

3188 (1996).<br />

[75] Y. Tomioka, A. Asamitsu, Y. Moritomo, H. Kuwahara, and Y. Tokura,<br />

Phys. Rev. Lett. 74, 5108 (1995).<br />

[76] P. W. Selwood, Magnetochemistry (Interscience, New York and London,<br />

1956).<br />

[77] M. M. Schieber, Experimental Magnetochemistry (Wiley, New York,<br />

1967).<br />

[78] S. Shoner and J. Kovacs, (unpublished).<br />

[79] T. Moriya, J. Magn. Magn. Mater. 100, 261-271 (1991).


186 References<br />

[80] E. C. Stoner, Proc. R. Soc. London Ser. A: 165, 372 (1938).<br />

[81] D. M. Edwards and E. P. Wohlfarth, Proc. R. Soc. London Ser. A: 303, 127-<br />

137 (1968).<br />

[82] E. C. Stoner, Acta Metallurgica 2, 259 (1954).<br />

[83] E. P. Wohlfarth, Phys. Lett. 3, 17 (1962).<br />

[84] T. Moriya, Spin Fluctuations in Itinerant Electron Magnetism (Springer<br />

Verlag, 1985).<br />

[85] T. Moriya, in Metallic Magnetism, Topics in Current Physics, edited by H.<br />

Capellman (Springer, Berlin, 1987), Vol. 42, p. 15-55.<br />

[86] G. G. Lonzarich and L. Taillefer, J. Phys. C: Solid State Phys. 18, 4339<br />

(1985).<br />

[87] K. Adachi, K. Sato, and M. Takeda, J. Phys. Soc. Jpn. 26, 631 (1969).<br />

[88] E. P. Wohlfarth, in Ferrmagnetic Materials, Vol. 1, p. 1-70.<br />

[89] W. Bao, C. Broholm, G. Aeppli, J. M. Honig, P. Metcalf, and S. F. Travino,<br />

Phys. Rev. B 54, 3726 (1996).<br />

[90] B. Widom, J. Chem. Phys. 43, 3898 (1965).<br />

[91] K. Makoshi and T. Moriya, J. Phys. Soc. Jpn. 38, 10 (1975).<br />

[92] R. Anders and W. von Hörsten, Phys. Dat. 20 (1984).<br />

[93] M. Seeger, H. Kronmuller, and H. J. Blythe, J. Magn. Magn. Mater. 139,<br />

312 (1995).<br />

[94] R. M. White and T. H. Geballe, in Solid State Physics, edited by H.<br />

Ehrenreich, F. Seitz and D. Turnbull (Academic, New York, 1979), Vol. 15.<br />

[95] F. Keffer, in Handbuch der Physik (Springer Verlag, Berlin, 1966), Vol. 18,<br />

p. 1.<br />

[96] G. S. Rushbrooke and P. J. Wood, Mol. Phys. 1, 257 (1958).<br />

[97] V. N. Smolyaninova, J. J. Hamilton, R. L. Greene, Y. M. Mukovskii, S. G.<br />

Karabashev, and A. M. Balbashov, Phys. Rev. B 55 (1997).<br />

[98] M. Itoh, I. Natori, S. Kubota, and K. Motoya, J. Phys. Soc. Jpn. 63, 1486<br />

(1994).


References 187<br />

[99] J. A. Mydosh, Spin Glasses (Taylor & Francis, London, 1993).<br />

[100] H. Y. Hwang, S.-W. Cheong, P. G. Radaelli, M. Marezio, and B. Batlogg,<br />

Phys. Rev. Lett. 75, 914 (1995).<br />

[101] A. H. Morrish, The Physical Principles of Magnetism (John Wiley &<br />

Sons, New York, 1965).<br />

[102] K. A. Moler, Thesis (Stanford University, 1995)<br />

[103] G. J. Snyder, R. Hiskes, S. DiCarolis, M. R. Beasley, and T. H. Geballe,<br />

Phys. Rev. B 53, 14434 (1996).<br />

[104] H. L. Ju, J. Gopalakrishnan, J. L. Peng, G. C. X. Qi. Li, T. Venkatesan, and<br />

R. L. Greene, Phys. Rev. B 51, 6143 (1995).<br />

[105] R. von Helmolt, J. Wecker, K. Samwer, and K. Bärner, J. Magn. Magn.<br />

Mater. 151, 411 (1995).<br />

[106] I. O. Troyanchuk, JETP 75, 132 (1992).<br />

[107] T. M. Perekalina, T. A. Sivokon', S. A. Cherezyan, and I. É. Lipin'ski,<br />

Sov. Phys. Solid State 31, 1524 (1989).<br />

[108] R. Mahendiran, A. K. Raychaudhuri, A. Chainani, D. D. Sarma, and S. B.<br />

Roy, Appl. Phys. Lett. 66, 233 (1995).<br />

[109] S. E. Lofland, S. M. Bhagat, H. L. Ju, G. C. Xiong, T. Venkatesan, and R. L.<br />

Greene, (1995).<br />

[110] M. Domínguez, S. M. Bhagat, S. E. Lofland, J. S. Ramachandran, G. C.<br />

Xiong, H. L. Ju, R. L. Greene, and T. Venkatesan, EuroPhysics 32, 349<br />

(1995).<br />

[111] M. Domínguez, S. E. Lofland, S. M. Bhagat, A. K. Raychaudhuri, H. L. Ju,<br />

T. Venkatesan, and R. L. Greene, Sol. State Comm. 97, 193 (1996).<br />

[112] K. Kubo and N. Ohata, J. Phys. Soc. Jpn. 33, 21 (1972).<br />

[113] A. Urushibara, Y. Moritomo, T. Arima, G. Kido, and Y. Tokura, Phys.<br />

Rev. B 51, 14103 (1995).<br />

[114] M. F. Hundley, M. Hawley, R. H. Heffner, Q. X. Jia, J. J. Neumeier, J.<br />

Tesmer, J. D. Thompson, and X. D. Wu, Appl. Phys. Lett. 67, 860 (1995).


188 References<br />

[115] M. F. Hundley, J. J. Neumeier, R. H. Heffner, Q. X. Jia, X. D. Wu, and J. D.<br />

Thompson, J. Appl. Phys. 79, 4535 (1995).<br />

[116] J. Z. Liu, I. C. Chang, S. Irons, P. Klavins, R. N. Shelton, K. Song, and S.<br />

R. Wasserman, Appl. Phys. Lett. 66, 3218 (1995).<br />

[117] I. Mannari, Prog. Theor. Phys. 22, 335 (1959).<br />

[118] I. A. Campell and A. Fert, in Ferromagnetic Materials, edited by E. P.<br />

Wholfarth (North Holland, New York, 1982), Vol. 3, Chap. 9.<br />

[119] K. Udea and T. Moriya, J. Phys. Soc. Jpn. 39, 605 (1975).<br />

[120] A. H. Thompson, Phys. Rev. Lett. 35, 1786 (1975).<br />

[121] Y. Tokura, A. Urushibara, Y. Moritomo, T. Arima, A. Asamitsu, G. Kido,<br />

and N. Furukawa, J. Phys. Soc. Jpn. 63, 3931 (1994).<br />

[122] Y. Tokura, A. Urushibara, Y. Moritomo, A. Asamitsu, Y. Tomioka, T.<br />

Arima, and G. Kido, J. Mater. Sci. Eng. B 31, 187 (1995).<br />

[123] N. F. Mott and E. A. Davis, Electronic Processes in non-Crystalline<br />

Materials (Clarendon Press, Oxford, 1971).<br />

[124] J. M. D. Coey, M. Viret, and L. Ranno, Phys. Rev. Lett. 75, 3910 (1995).<br />

[125] M. E. Fisher and J. S. Langer, Phys. Rev. Lett. 20, 665 (1968).<br />

[126] S. C. Gausepohl, M. Lee, K. Char, and C. B. Eom, Phys. Rev. B 52, 3459<br />

(1995).<br />

[127] P. G. deGennes and J. Friedel, J. Phys. Chem. Solids 4, 71-77 (1958).<br />

[128] C. Haas, Phys. Rev. 168, 531 (1968).<br />

[129] Methfessel and Mattis, in Handbuch der Physik (Springer Verlag, Berlin,<br />

1968), Vol. 18, p. 389.<br />

[130] S. von Molnár and S. Methfessel, J. Appl. Phys. 38, 959 (1967).<br />

[131] S. von Molnár, I. Terry, T. Penney, and P. Becla, Physica B 197, 151 (1994).<br />

[132] V. J. Folen and G. H. Stauss, in Landolt-Börnstein Tabellen (Springer,<br />

Berlin, 1970), Vol. III/4b.<br />

[133] M. Böhm and A. Scharmann, in Landolt-Börnstein Tabellen, edited by<br />

O. Madelung (Springer, Berlin, 1982), Vol. III/17h.


References 189<br />

[134] H. von Philipsborn, L. Treitinger, and M. Rubenstein, in Landolt-<br />

Börnstein Tabellen (Springer, Berlin, 1978), Vol. III/12b.<br />

[135] M. R. Oliver, J. O. Dimmock, A. L. McWhorter, and T. B. Reed, Phys.<br />

Rev. B 5, 1078 (1972).<br />

[136] J. M. DeTeresa, M. R. Ibarra, J. Garcia, J. Blasco, C. Ritter, P. A. Algarabel,<br />

C. Marquina, and A. delMoral, Phys. Rev. Lett. 76, 3392 (1996).<br />

[137] G. J. Snyder, C. Booth, F. Bridges, R. Hiskes, S. DiCarolis, M. R. Beasley,<br />

and T. H. Geballe, Phys. Rev. B (1997).<br />

[138] R. Pauthenet, J. Appl. Phys. 29, 253 (1958).<br />

[139] H. Yoshizawa, H. Kawano, Y. Tomioka, and Y. Tokura, Phys. Rev. B 52,<br />

13145 (1995).<br />

[140] I. S. Jacobs, J. Phys. Chem. Solids 11, 1 (1959).<br />

[141] I. S. Jacobs, J. Phys. Chem. Solids 15, 54 (1960).<br />

[142] A. S. Yermolenko, Phys. Met. Metallogr. 50(4), 57 (1980).<br />

[143] Z. Jirak, S. Krupicka, Z. Simsa, M. Dlouha, and S. Vratislav, J. Magn.<br />

Magn. Mater. 53, 153 (1985).<br />

[144] J. J. Melero, F. Bartolome, J. Bartolome, and R. Burriel, J. Magn. Magn.<br />

Mater. (unpublished).<br />

[145] Z. Jirak, S. Krupicka, V. Nekvasil, E. Pollert, G. Villeneuve, and F.<br />

Zounova, J. Magn. Magn. Mater. 15-18, 519 (1980).<br />

[146] D. C. Worledge, G. J. Snyder, R. Hiskes, S. DiCarolis, M. R. Beasley, and<br />

T. H. Geballe, J. Appl. Phys. 80, 5158 (1996).<br />

[147] P. Shuk, L. Tichonova, and U. Guth, Solid State Ionics 68, 177 (1994).<br />

[148] H. Taguchi, M. Nagao, and M. Shimada, J. Solid State Chem. 82, 8 (1989).<br />

[149] V. Ambegaokar, B. I. Halperin, and J. S. Langer, Phys. Rev. B 4, 2612<br />

(1971).<br />

[150] G. J. Snyder, R. Hiskes, S. DiCarolis, M. R. Beasley, and T. H. Geballe,<br />

Appl. Phys. Lett. 69, 4254 (1996).


190 References<br />

[151] S. J. L. Billinge, R. G. DiFrancesco, G. H. Kwei, J. J. Neumeier, and J. D.<br />

Thompson, Phys. Rev. Lett. 77, 715 (1996).<br />

[152] R. P. Sharma, G. C. Xiong, C. Kwon, R. Ramesh, R. L. Greene, and T.<br />

Venkatesan, Phys. Rev. B 54, 10014 (1996).<br />

[153] P. G. Radaelli, M. Marezio, H. Y. Hwang, S.-W. Cheong, and B. Batlogg,<br />

Phys. Rev. B 54, 8992 (1996).<br />

[154] C. H. Booth, F. Bridges, G. J. Snyder, and T. H. Geballe, Phys. Rev. B<br />

(1996).<br />

[155] C. W. Searle and S. T. Wang, Can. J. Phys. 48, 2023 (1970).<br />

[156] Y. X. Jia, L. Lu, K. Khazeni, V. H. Crespi, A. Zettl, and M. L. Cohen, Phys.<br />

Rev. B 52, 9147 (1995).<br />

[157] L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media<br />

(Pergamon, Oxford, 1984).<br />

[158] J. O'Donnell, M. Onellion, M. S. Rzchowski, J. N. Eckstein, and I.<br />

Bozovic, Phys. Rev. B (1996).<br />

[159] J. N. Eckstein, I. Bozovic, J. O'Donnell, M. Onellion, and M. S.<br />

Rzchowski, Appl. Phys. Lett. (1996).<br />

[160] G. Bergmann, Phys. Today August 1979, 25 (1979).<br />

[161] J. E. Nunez-Regueiro and A. M. Kadin, Appl. Phys. Lett. 68, 2747 (1996).<br />

[162] J. Inoue and S. Maekawa, Phys. Rev. Lett. 74, 3407 (1995).<br />

[163] N. Furukawa, J. Phys. Soc. Jpn. 64, 3164 (1995).<br />

[164] N. Furukawa, J. Phys. Soc. Jpn. 64, 2734 (1995).<br />

[165] S. Yamaguchi, H. Taniguchi, H. Takagi, T. Arima, and Y. Tokura, J. Phys.<br />

Soc. Jpn. 64, 1885 (1995).<br />

[166] M. Jaime, M. B. Salamon, M. Rubinstein, R. E. Treece, J. S. Horwitz, and<br />

D. B. Chrisey, Phys. Rev. B 54, 11914 (1996).<br />

[167] R. H. Heffner, L. P. Lee, M. F. Hundley, J. J. Neumeier, G. M. Luke, K.<br />

Kojima, B. Nachumi, Y. J. Uemura, D. E. MacLaughlin, and S.-W.<br />

Cheong, Phys. Rev. Lett. 77, 1869 (1996).


References 191<br />

[168] J. F. Mitchell, D. N. Argyriou, C. D. Potter, J. D. Jorgensen, D. G. Hinks,<br />

and S. D. Bader, Mat. Res. Soc. Symp. Proc. 453 (1996).<br />

[169] J. F. Mitchell, (unpublished).<br />

[170] Lynn, Phys. Rev. Lett. 76, 4046 (1996).<br />

[171] P. G. Radaelli, D. E. Cox, M. Marezio, S.-W. Cheong, P. E. Schiffer, and A.<br />

P. Ramirez, Phys. Rev. Lett. 75, 4488 (1995).<br />

[172] J. L. García-Muñoz, M. Suaaidi, J. Fontcuberta, and J. Rodríguez-<br />

Carvajal, Phys. Rev. B 55, 34 (1997).<br />

[173] Y. Noro and S. Miyahara, J. Phys. Soc. Jpn. 27, 518 (1969).<br />

[174] M. Shikano, T.-K. Huang, Y. Inaguma, M. Itoh, and T. Nakamura, Sol.<br />

State Comm. 90, 115 (1994).<br />

[175] J. J. Neumeier, A. L. Cornelius, and J. S. Schilling, Physica B 198, 324<br />

(1994).<br />

[176] P. B. Allen, H. Berger, O. Chauvet, L. Forro, T. Jarlborg, A. Junod, B.<br />

Revaz, and G. Santi, Phys. Rev. B 53, 84393 (1996).<br />

[177] C. W. Jones, P. D. Battle, and P. Lightfoot, Acta Ctyst. C45, 365 (1989).<br />

[178] J. M. Longo, P. M. Raccah, and J. B. Goodenough, J. Appl. Phys. 39, 1327<br />

(1968).<br />

[179] A. Callaghan, C. W. Moeller, and R. Ward, Inorg. Chem. 5, 1572 (1966).<br />

[180] D. J. Singh, J. Appl. Phys. 79 (1996).<br />

[181] A. Kanbayasi, J. Phys. Soc. Jpn. 41, 1876 (1976).<br />

[182] A. Kanbayasi, J. Phys. Soc. Jpn. 41, 1879 (1976).<br />

[183] A. Kanbayasi, J. Phys. Soc. Jpn. 44, 89 (1978).<br />

[184] W. Bensch, H. W. Schmalle, and A. Reller, Solid State Ionics 43, 171<br />

(1990).<br />

[185] B. D. Cullity, Introduction to Magnetic Materials (Addison-Wesley,<br />

Reading, Massachusetts, 1972).<br />

[186] L. Klein, J. S. Dodge, T. H. Geballe, A. Kapitulnik, A. F. Marshall, L.<br />

Antognazza, and K. Char, Appl. Phys. Lett. 66, 2427 (1995).


192 References<br />

[187] L. Klein, (unpublished).<br />

[188] A. H. Bobeck, Magnetic Bubbles (Elsevier, New York, 1975).<br />

[189] R. Pauthenet, in High Field Magnetism, edited by M. Date (North-<br />

Holland, 1983), p. 77.<br />

[190] B. E. Argyle, S. H. Charap, and E. W. Pugh, Phys. Rev. 132, 2051 (1963).<br />

[191] A. T. Aldred and P. H. Froehle, Int. J. Magn. 2, 195-203 (1972).<br />

[192] A. T. Aldred, Phys. Rev. B 11, 2597 (1975).<br />

[193] E. P. Wohlfarth and P. F. DeChatel, Physica 48, 477-485 (1970).<br />

[194] J. Takeuchi and Y. Masuda, J. Phys. Soc. Jpn. 46, 468 (1979).<br />

[195] S. M. Hayden and C. C. Paulsen, Physica B 166, 201 (1990).<br />

[196] J. J. Neumeier, private communication.<br />

[197] P. M. Levy, Phys. Rev. 170, 595 (1968).<br />

[198] P. M. Levy and D. P. Landau, J. Appl. Phys. 39, 1128 (1968).<br />

[199] K. Ikeda and K. A. Gschneider, J. Magn. Magn. Mater. 30, 273 (1983).<br />

[200] J. A. Rayne and W. R. G. Kemp, Phil. Mag. 1, 918 (1956).<br />

[201] J. A. Rayne and B. S. Chandrasekhar, Phys. Rev. 122, 1714 (1961).<br />

[202] M. Dixon, F. E. Hoare, T. M. Holden, and D. E. Moody, Proc. R. Soc.<br />

London Ser. A: 285, 561 (1965).<br />

[203] S. Shioda, Y. Takahashi, and T. Moriya, J. Phys. Soc. Jpn. 57, 3146 (1988).<br />

[204] P. Hertel, J. Appel, and D. Fay, Phys. Rev. B 22, 534 (1980).<br />

[205] K. Ikeda and M. Yoshizawa, Physica B 165-166, 203 (1990).

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!